parameter identifiability, constraint, and equifinality in

26
Ecological Applications, 19(3), 2009, p. 571–574 Ó 2009 by the Ecological Society of America Parameter identifiability, constraint, and equifinality in data assimilation with ecosystem models YIQI LUO, 1,3 ENSHENG WENG, 1 XIAOWEN WU, 1 CHAO GAO, 1 XUHUI ZHOU, 1 AND LI ZHANG 2 1 Department of Botany and Microbiology, University of Oklahoma, Norman, Oklahoma 73019 USA 2 Institute of Geographic Science and Natural Resources Research, Chinese Academy Sciences, Beijing 100101, China One of the most desirable goals of scientific endeavor is to discover laws or principles behind ‘‘mystified’’ phenomena. A cherished example is the discovery of the law of universal gravitation by Isaac Newton, which can precisely describe falling of an apple from a tree and predict the existence of Neptune. Scientists pursue mechanistic understanding of natural phenomena in an attempt to develop relatively simple equations with a small number of parameters to describe patterns in nature and to predict changes in the future. In this context, uncertainty had been considered to be incom- patible with science (Klir 2006). Not until the early 20th century was the notion gradually changed when physicists studied the behavior of matter and energy on the scale of atoms and subatomic particles in quantum mechanics. In 1927, Heisenberg observed that the electron could not be considered as in an exact location, but rather in points of probable location in its orbital, which can be described by a probability distribution (Heisenberg 1958). Quantum mechanics lets scientists realize that inherent uncertainty exists in nature and is an unavoidable and essential property of most systems. Since then, scientists have developed methods to analyze and describe uncertainty. Ecosystem ecologists have recently directed attention to studying uncertainty in ecosystem processes. The Bayesian paradigm allows ecologists to generate a posteriori probabilistic density functions (PDF) for parameters of ecosystem models by assimilating a priori PDFs and measurements (Dowd and Meyer 2003). Xu et al. (2006), for example, evaluated uncertainty in parameter estimation and projected carbon sinks by a Bayesian framework using six data sets and a terrestrial ecosystem (TECO) model. The Bayesian framework has been applied to assimilation of eddy-flux data into simplified photosynthesis and evapotranspiration model (SIPNET) to evaluate information content of the net ecosystem exchange (NEE) observations for constraints of process parameters (e.g., Braswell et al. 2005) and to partition NEE into its component fluxes (Sacks et al. 2006). Verstraeten et al. (2008) evaluate error propaga- tion and uncertainty of evaporation, soil moisture content, and net ecosystem productivity with remotely sensed data assimilation. Nevertheless, uncertainty in data assimilation with ecosystem models has not been systematically explored. Cressie et al. (2009) proposed a general framework to account for multiple sources of uncertainty in measure- ments, in sampling, in specification of the process, in parameters, and in initial and boundary conditions. They proposed to separate the multiple sources of uncertainty using a conditional-probabilistic approach. With this approach, ecologists need to build a hierarch- ical statistical model based on the Bayesian theorem, and to use Markov chain Monte Carlos (MCMC) techniques for sampling before probability distributions of interested parameters or projected state variables can be obtained for quantification of uncertainty. It is an elegant framework for quantifying uncertainties in the parameters and processes of ecological models. At the core of uncertainty analysis is parameter identifiability. When parameters can be constrained by a set of data with a given model structure, we can identify maximum likelihood values of the parameters and then those parameters are identifiable. Conversely, there is an issue of equifinality in data assimilation (Beven 2006) that different models, or different parameter values of the same model, may fit data equally well without the ability to distinguish which models or parameter values are better than others. Thus, the issue of identifiability is reflected by parameter constraint and equifinality. This essay first reviews the current status of our knowledge on parameter identifiability and then discusses major factors that influence it. To enrich discussion, we use examples in ecosystem ecology that are different from the one on population dynamics of harbor seals in Cressie et al. (2009). CURRENT STATUS Data assimilation is in the infant stage in ecosystem ecology but gradually is becoming a more active research subject due to increased data availability from observational networks. Uncertainty analysis with data assimilation has been conducted in a limited number of studies in the past few years (Xu et al. 2006, Verstraeten Manuscript received 20 March 2008; revised 17 June 2008; accepted 30 June 2008. Corresponding Editor: N. T. Hobbs. For reprints of this Forum, see footnote 1, p. 551. 3 E-mail: [email protected] 571 April 2009 HIERARCHICAL MODELS IN ECOLOGY

Upload: others

Post on 18-Dec-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Ecological Applications, 19(3), 2009, p. 571–574� 2009 by the Ecological Society of America

Parameter identifiability, constraint, and equifinality in dataassimilation with ecosystem models

YIQI LUO,1,3 ENSHENG WENG,1 XIAOWEN WU,1 CHAO GAO,1 XUHUI ZHOU,1 AND LI ZHANG2

1Department of Botany and Microbiology, University of Oklahoma, Norman, Oklahoma 73019 USA2Institute of Geographic Science and Natural Resources Research, Chinese Academy Sciences, Beijing 100101, China

One of the most desirable goals of scientific endeavor

is to discover laws or principles behind ‘‘mystified’’

phenomena. A cherished example is the discovery of the

law of universal gravitation by Isaac Newton, which can

precisely describe falling of an apple from a tree and

predict the existence of Neptune. Scientists pursue

mechanistic understanding of natural phenomena in an

attempt to develop relatively simple equations with a

small number of parameters to describe patterns in

nature and to predict changes in the future. In this

context, uncertainty had been considered to be incom-

patible with science (Klir 2006). Not until the early 20th

century was the notion gradually changed when

physicists studied the behavior of matter and energy

on the scale of atoms and subatomic particles in

quantum mechanics. In 1927, Heisenberg observed that

the electron could not be considered as in an exact

location, but rather in points of probable location in its

orbital, which can be described by a probability

distribution (Heisenberg 1958). Quantum mechanics lets

scientists realize that inherent uncertainty exists in

nature and is an unavoidable and essential property of

most systems. Since then, scientists have developed

methods to analyze and describe uncertainty.

Ecosystem ecologists have recently directed attention

to studying uncertainty in ecosystem processes. The

Bayesian paradigm allows ecologists to generate a

posteriori probabilistic density functions (PDF) for

parameters of ecosystem models by assimilating a priori

PDFs and measurements (Dowd and Meyer 2003). Xu

et al. (2006), for example, evaluated uncertainty in

parameter estimation and projected carbon sinks by a

Bayesian framework using six data sets and a terrestrial

ecosystem (TECO) model. The Bayesian framework has

been applied to assimilation of eddy-flux data into

simplified photosynthesis and evapotranspiration model

(SIPNET) to evaluate information content of the net

ecosystem exchange (NEE) observations for constraints

of process parameters (e.g., Braswell et al. 2005) and to

partition NEE into its component fluxes (Sacks et al.

2006). Verstraeten et al. (2008) evaluate error propaga-

tion and uncertainty of evaporation, soil moisture

content, and net ecosystem productivity with remotely

sensed data assimilation. Nevertheless, uncertainty in

data assimilation with ecosystem models has not been

systematically explored.

Cressie et al. (2009) proposed a general framework to

account for multiple sources of uncertainty in measure-

ments, in sampling, in specification of the process, in

parameters, and in initial and boundary conditions.

They proposed to separate the multiple sources of

uncertainty using a conditional-probabilistic approach.

With this approach, ecologists need to build a hierarch-

ical statistical model based on the Bayesian theorem,

and to use Markov chain Monte Carlos (MCMC)

techniques for sampling before probability distributions

of interested parameters or projected state variables can

be obtained for quantification of uncertainty. It is an

elegant framework for quantifying uncertainties in the

parameters and processes of ecological models.

At the core of uncertainty analysis is parameter

identifiability. When parameters can be constrained by a

set of data with a given model structure, we can identify

maximum likelihood values of the parameters and then

those parameters are identifiable. Conversely, there is an

issue of equifinality in data assimilation (Beven 2006)

that different models, or different parameter values of

the same model, may fit data equally well without the

ability to distinguish which models or parameter values

are better than others. Thus, the issue of identifiability is

reflected by parameter constraint and equifinality. This

essay first reviews the current status of our knowledge

on parameter identifiability and then discusses major

factors that influence it. To enrich discussion, we use

examples in ecosystem ecology that are different from

the one on population dynamics of harbor seals in

Cressie et al. (2009).

CURRENT STATUS

Data assimilation is in the infant stage in ecosystem

ecology but gradually is becoming a more active

research subject due to increased data availability from

observational networks. Uncertainty analysis with data

assimilation has been conducted in a limited number of

studies in the past few years (Xu et al. 2006, Verstraeten

Manuscript received 20 March 2008; revised 17 June 2008;accepted 30 June 2008. Corresponding Editor: N. T. Hobbs. Forreprints of this Forum, see footnote 1, p. 551.

3 E-mail: [email protected]

571

April 2009 HIERARCHICAL MODELS IN ECOLOGY

et al. 2008). However, parameter identifiability in

association with uncertainty analysis has not been much

explored at all. That is not because it is not an issue for

data assimilation with ecosystem models but because

this issue has not been well formulated to rise to the top

of the research agenda in the community.

In reality, the number of identifiable parameters in

any process-based ecosystem models is extremely low,

particularly when only NEE data from eddy-flux

networks are used in data assimilation. For example,

Wang et al. (2001) showed that only a maximum of three

or four parameters could be determined independently

from the CO2 flux observation. Posterior standard

deviations were reduced in seven out of 14 BETHY C4

plant model parameters and in five out of 23 BETHY C3

plant model parameters relative to prior standard

deviation of parameters (Knorr and Kattge 2005). When

six data sets of soil respiration, woody biomass, foliage

biomass, litter fall, and soil carbon content were used for

parameter estimation, four out of seven carbon transfer

coefficients at ambient CO2 and three at elevated CO2

can be constrained (Xu et al. 2006). The parameter

identifiability is a ubiquitous issue in data assimilation

with ecosystem models.

Most of the published studies on data assimilation

with ecosystem models avoid the issue of parameter

identifiability by using simplified ecosystem models with

a very limited set of parameters and/or prescribing

many of the parameters in the models. Williams et al.

(2005), for example, used a simple carbon box model

with nine parameters and five initial values of carbon

pools to be estimated from eddy-flux data using

ensemble Kalman Filter. Braswell et al. (2005) used a

three-carbon-pool model and estimated 23 parameters

while initial values of the leaf carbon pool and soil

moisture content were prescribed. Luo and his col-

leagues prescribed partitioning coefficients of photo-

synthates into plant pools, initial values of pool sizes,

and parameters that describe carbon flows into receiving

pools in the inverse analysis with six data sets from a

forest CO2 experiment (Luo et al. 2003, Xu et al. 2006).

They only estimated seven transfer coefficients from the

plant and soil carbon pools based on the rationale that

the coefficients are among the most important param-

eters in determining ecosystem carbon sequestration

(Luo et al. 2003). None of the studies has used

comprehensive ecosystem models for data assimilation

because of the difficulty in identifying hundreds of

parameters against limited sets of data. In short, the

issue of parameter identifiability in data assimilation has

not been explicitly addressed although equifinality is a

serious problem in ecosystem modeling (Medlyn et al.

2005). A Bayesian framework described by Cressie et al.

(2009) is useful for examining causes of equifinality or

parameter identifiability as related to uncertainty in

measurements, data availability, specification of model

structure, and optimization methods.

FACTORS THAT INFLUENCE PARAMETER IDENTIFIABILITY

IN ECOSYSTEM MODELS

Many factors influence parameter identifiability in

ecosystem models, such as data availability, modelstructures, optimization methods, initial values, boun-

dary conditions, ranges, and patterns of priors. Sincenone of the factors has been explicitly examined in the

literature of ecosystem research, the following discussionis mainly derived from our own studies on data

availability and compatibility, model structures (E. S.Weng and Y. Q. Luo, unpublished data), and optimiza-

tion methods.Data availability and compatibility.—It is a common

sense that we have to collect relevant data to address aspecific scientific issue. For example, we measure

mineralization rates to examine nitrogen availability.To characterize ecosystem carbon cycling, we have to

collect carbon-related data, such as photosynthesis,respiration, plant growth, and soil carbon content andfluxes.

Recently, eddy-flux networks have been set upworldwide and offer extensive data sets that record net

ecosystem exchange of CO2 (NEE) between the atmos-phere and ecosystems. These data sets hold great

promises for understanding the processes of ecosystemcarbon cycling. Several studies have been conducted on

NEE data assimilation to improve predictive under-standing of ecosystem carbon processes (Braswell et al.

2005, Williams et al. 2005, Wang et al. 2007). However,it has not been carefully evaluated how different data

sets constrain different model parameters.L. Zhang, Y. Q. Luo, G. R. Yu, and L. M. Zhang

(unpublished manuscript) recently conducted a study toevaluate parameter identifiability using NEE and bio-

metric data in three forest ecosystems in China.Biometric data in their study include observed values

of foliage biomass, fine-root biomass, woody biomass,litter fall, soil organic carbon, and soil respiration. Threeexperiments of data assimilation were conducted to

estimate carbon transfer coefficients using biometricdata only, NEE data only, or both the biometric and

NEE data.Biometric data were effective in constraining carbon

transfer coefficients from plants pools in leaves, roots,and wood, whereas NEE data strongly constrained

carbon transfer coefficients from litter, microbial, andslow soil pools. It indicated that measurements of

foliage, fine-root, and woody biomass provided infor-mation on C transfer from plant to litter pools. NEE

data provided the information of C transfer amonglitter, microbial biomass, and SOM pools, from which

CO2 was released. In addition, the NEE data set has amuch larger sample size than those of biometric data

sets. As a result, biometric data sets provided lessinformation than the NEE data set in constrainingparameters. When NEE data without gap-filled points

were used, the transfer coefficients from litter, microbial,and slow soil carbon pools were much less constrained.

FORUM572Ecological Applications

Vol. 19, No. 3

Parameter identifiability is also dependent on rele-

vance of data. In a data assimilation study with a

mechanistic canopy photosynthesis model, two activa-

tion energy parameters for carboxylation and oxygen-

ation cannot be constrained by NEE data. To constrain

those parameters, we need measurements of enzyme

kinetics along a temperature gradient. In general,

different kinds of data constrain different parameters.

However, magnitudes of measurement errors were not

found to considerably influence parameter identifiabil-

ity. We also need to evaluate information content of

different length, frequency, and quality of measurement

data in constraining different parameters.

Model structures.—Model structure is a main source

of uncertainty (Chatfield 1995). We have a tendency to

incorporate more and more processes into models to

improve fitness between simulated and observed data.

Complicated models may integrate more process knowl-

edge but make more parameters less identifiable given

certain data sets. We conducted a data assimilation

study to evaluate parameter identifiability when carbon

pools vary from three to eight among models. The eight-

pool model has foliage, woody, roots, metabolic litter,

structural litter, microbial, slow soil carbon, and passive

soil carbon pools. We lumped together plant pools, litter

pools, and soil pools in different combinations to

construct four other models with pools ranging from

three to seven. Nine data sets (i.e., foliage biomass,

woody biomass, fine-root biomass, microbial biomass,

litter fall, forest floor carbon, organic soil carbon,

mineral soil carbon, and soil respiration) from the Duke

CO2 experiment site (North Carolina. USA) constrain

the three-pool model best. Transfer coefficients from

passive soil carbon pool and microbial pool could not be

identified in seven or eight pool models. Thus, intro-

duction of additional processes in modeling may reduce

parameter identifiability.

Optimization methods.—Many papers and books have

been published on methods for local and global

optimization. We have evaluated different optimization

methods for parameter identifiability by comparing

conventional and conditional Bayesian inversions. The

conditional Bayesian inversion sequentially identifies

constrained parameters in several steps. In each step,

constrained parameters were obtained. Their maximum

likelihood estimators (MLE) were used as priors to fix

the parameter values at MLE in the next step of

Bayesian inversion with decreased parameter dimen-

sionality. Conditional inversion was repeated until there

was no more parameter to be constrained. This method

was applied with a physiologically based ecosystem

model to hourly NEE measurements at Harvard Forest

(Petersham, Massachusetts, USA). The conventional

inversion method constrained six out of 16 parameters

in the model while the conditional inversion method

constrained 14 parameters after six steps.

The conditional inversion resulted in more con-

strained parameters for two reasons. First, parameter

dimensionality decreased during conditional Bayesian

inversion, resulting in less difficulty in finding the

optimized parameter values. Second, parameter identi-

ifiability is altered due to increased differentiability

against NEE data once some other parameter values are

fixed within the steps of conditional Bayesian inversion.

In summary, parameter identifiability is a critical but

complex issue that needs to be carefully addressed in the

future. The hierarchical Bayesian modeling described by

Cressie et al. (2009) is a promising approach to explore

different factors in influencing parameter identifiability

or equifinality.

ACKNOWLEDGMENTS

This research was financially supported by National ScienceFoundation (NSF) under DEB 0444518 and DEB 0714142, andby the Office of Science (BER), Department of Energy, GrantNo. DE-FG02-006ER64319.

LITERATURE CITED

Beven, K. 2006. A manifesto for the equifinality thesis. Journalof Hydrology 320:18–36.

Braswell, B. H., W. J. Sacks, E. Linder, and D. S. Schimel.2005. Estimating diurnal to annual ecosystem parameters bysynthesis of a carbon flux model with eddy covariance netecosystem exchange observations. Global Change Biology11:335–355.

Chatfield, C. 1995. Model uncertainty, data mining andstatistical-inference. Journal of the Royal Statistical SocietySeries A: Statistics in Society 158:419–466.

Cressie, N., C. A. Calder, J. S. Clark, J. M. Ver Hoef, and C. K.Wikle. 2009. Accounting for uncertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19:553–570.

Dowd, M., and R. Meyer. 2003. A Bayesian approach to theecosystem inverse problem. Ecological Modelling 168:39–55.

Heisenberg, W. 1958. Physics and philosophy; the revolution inmodern science. Harper, New York, New York, USA.

Klir, G. J. 2006. Uncertainty and information. John Wiley andSons, Hoboken, New Jersey, USA.

Knorr, W., and J. Kattge. 2005. Inversion of terrestrialecosystem model parameter values against eddy covariancemeasurements by Monte Carlo sampling. Global ChangeBiology 11:1333–1351.

Luo, Y. Q., L. W. White, J. G. Canadell, E. H. DeLucia, D. S.Ellsworth, A. Finzi, J. Lichter, and W. H. Schlesinger. 2003.Sustainability of terrestrial carbon sequestration: a case studyin Duke Forest with inversion approach. Global Biogeo-chemical Cycles 17. [doi: 10.1029/2002GB001923]

Medlyn, B. E., A. P. Robinson, R. Clement, and R. E.McMurtrie. 2005. On the validation of models of forest CO2

exchange using eddy covariance data: some perils andpitfalls. Tree Physiology 25:839–857.

Sacks, W. J., D. S. Schimel, R. K. Monson, and B. H. Braswell.2006. Model-data synthesis of diurnal and seasonal CO2

fluxes at Niwot Ridge, Colorado. Global Change Biology 12:240–259.

Verstraeten, W. W., F. Veroustraete, W. Heyns, T. Van Roey,and J. Feyen. 2008. On uncertainties in carbon flux modellingand remotely sensed data assimilation: the Brasschaat pixelcase. Advances in Space Research 41:20–35.

Wang, Y. P., D. Baldocchi, R. Leuning, E. Falge, and T.Vesala. 2007. Estimating parameters in a land-surface modelby applying nonlinear inversion to eddy covariance fluxmeasurements from eight FLUXNET sites. Global ChangeBiology 13:652–670.

Wang, Y. P., R. Leuning, H. A. Cleugh, and P. A. Coppin.2001. Parameter estimation in surface exchange models using

April 2009 573HIERARCHICAL MODELS IN ECOLOGY

nonlinear inversion: how many parameters can we estimate

and which measurements are most useful? Global Change

Biology 7:495–510.

Williams, M., P. A. Schwarz, B. E. Law, J. Irvine, and M. R.

Kurpius. 2005. An improved analysis of forest carbon

dynamics using data assimilation. Global Change Biology11:89–105.

Xu, T., L. White, D. Hui, and Y. Luo. 2006. Probabilisticinversion of a terrestrial ecosystem model: analysis of un-certainty in parameter estimation andmodel prediction.GlobalBiogeochemical Cycles 20. [doi: 10.1029/2005GB002468]

________________________________

Ecological Applications, 19(3), 2009, pp. 574–577� 2009 by the Ecological Society of America

The importance of accounting for spatial and temporal correlationin analyses of ecological data

JENNIFER A. HOETING1

Department of Statistics, Colorado State University, Fort Collins, Colorado 80523-1877 USA

I congratulate Cressie, Calder, Clark, Ver Hoef, and

Wikle for their insightful overview on hierarchical

modeling in ecology. These authors are all experts in

this field and have contributed to many of the advances

that have been made in the field of ecological modeling.

Spatial and temporal correlation is a major theme in

Cressie et al. (2009). Below I expand on these ideas,

discussing some of the advantages and disadvantages of

accounting for spatial and/or temporal correlation in

analyses of ecological data. While much of the focus in

this discussion is on spatial statistical models, similar

problems occur when temporal or spatiotemporal

correlation is ignored.

AN EXAMPLE OF A STATISTICAL MODEL THAT ACCOUNTS

FOR SPATIAL CORRELATION

Unless data are observed within a very specific

experimental design, ecological data are often corre-

lated. As an example, consider the problem of estimating

stream sulfate concentrations in the eastern United

States. We consider data collected as part of the EPA’s

Environmental Monitoring and Assessment Program

(EMAP). The sample sites were mainly located in

Pennsylvania, West Virginia, Maryland, and Virginia.

For more details about this example and the issues

described below, see Irvine et al. (2007).

The response Y ¼ [Y1, . . . , Yn]0 is stream sulfate

concentration at each of the n stream sites. In this

simplified example we consider four predictors [X1, . . . ,

X4] which are geographic information system (GIS)

derived covariates of the percentage of landscape

covered by forest, agriculture, urban, and mining within

the watershed above each stream site. To investigate the

relationship between these covariates and the response,

we might first consider a standard multiple regression

model for i¼ 1, . . . , n given by

Yi ¼ b0 þ b1Xi1 þ b2Xi2 þ b3Xi3 þ b4Xi4 þ ei ð1Þ

where Xij is the ith observation from the jth predictor

and the error terms ei are independent normally

distributed random variables with mean 0 and variance

r2. If we follow through with this analysis, we might find

that various predictors have statistically significant

effects, and we might provide some estimate of the

variance, r2. We might also undertake some form of

model selection to determine which covariates best

explain observed patterns of stream sulfate concentra-

tion. However, what if, after accounting for all available

covariates, the errors are not independent? What if there

is remaining spatial correlation, such that observations

that are in close proximity in space are related, and that

these predictors haven’t fully accounted for such

correlation? In this case, we can learn a lot from

modeling the nonindependent errors.

As an alternative to a multiple regression model with

independent errors, we might consider a spatial regres-

sion model where the errors are assumed to be spatially

correlated. In this case, we could use the model in Eq. 1,

but now would assume that the errors, ei, for stream

sites close to one another are more similar than the

errors for stream sites that are far apart. In mathemat-

ical terms, the independent error model assumes e ;

N(0, r2I) where I is the n 3 n identity matrix and 0 is a

vector of length n, and the spatial regression model

assumes e ; N(0, r2R) where R is an n 3 n correlation

matrix. This model goes by many names in the literature

and is a type of general (or generalized) linear model.

For our stream sulfate problem, we might adopt a model

for the covariance matrix R (see e.g., Schabenberger and

Manuscript received 2May 2008; revised 28May; accepted 30June 2008. Corresponding Editor: N. T. Hobbs. For reprints ofthis Forum, see footnote 1, p. 551.

1 E-mail: [email protected]

FORUM574Ecological Applications

Vol. 19, No. 3

Gotway 2005) and conclude that observations that are

less than 200 km apart are related. This might lead us to

think about the biological and physical processes that

could lead to this correlation; such research avenues

might suggest additional covariates that should be

included in our model.

DISADVANTAGES OF IGNORING SPATIAL CORRELATION

What could go wrong if we use the independent error

model (Eq. 1) instead of a spatial regression model when

the latter is the appropriate model? Plenty. There is a

long history of research that demonstrates the many

disadvantages of ignoring spatial correlation. Some of

the highlights and relations to ecology are described here.

One key issue is sample size. If an independent-error

model is adopted (Eq. 1) but the model errors are not

independent, then the ‘‘effective sample size’’ will be

smaller than the number of observations collected

(Schabenberger and Gotway 2005:32). Effective sample

size decreases as the correlation between observations

increases. If an independent-error model is adopted but

the data are correlated, standard errors can be under-

estimated. For example, when the independent-error

model is used and maximum likelihood estimates of the

regression coefficients b1, . . . , b4 in Eq. 1 are obtained,

the parameter estimates will be unbiased but the

standard errors of these estimates can be too small

(Schabenberg and Gotway 2005:324). In ecology this

underestimate of uncertainty can be critical: a covariate

may be deemed to be important only because an

inappropriate model is selected.

In the area of model selection, Hoeting et al. (2006)

showed that ignoring spatial correlation when selecting

covariates for inclusion in regression models can lead to

the exclusion of relevant covariates in the model.

Ignoring spatial correlation in model selection can also

lead to higher prediction errors for estimation of the

response.

The drawbacks described above were based on

research for non-Bayesian spatial modeling. In addition

to those drawbacks, non-Bayesian spatial models can

lead to underestimation of uncertainty. For example,

traditional estimation methods for the spatial regression

model assume that the covariance matrix R is fixed even

when the parameters in the model for R are estimated.

This leads to estimates of standard errors that do not

account for the uncertainty in all parameters.

Spatial correlation also plays a factor in sampling

design. Cressie et al. (2009) made an important point

that hierarchical models allow for direct incorporation

of the sampling design in the modeling. The advantages

of a sound sampling design cannot be overemphasized.

Too many ecological studies involve sites selected for

convenience. When the goal of an analysis is to provide

a map or some other inference across a sampling area,

then additional considerations should be made when

designing the study. It has been shown in a number of

contexts that a cluster sampling design is appropriate for

spatially correlated data (e.g., Zimmerman 2006, Irvine

et al. 2007, Ritter and Leecaster 2007). A cluster design

includes some observations observed at close distances

as well as sampling coverage over the entire sampling

area. Xia et al. (2006) proposed methodology that

produces an optimal design for spatially correlated data

where the optimization and subsequent design depends

on the goals of the study. For example, a design which

emphasizes accurate estimation of the regression coef-

ficients in Eq. 1 will be different than a design which

emphasizes accurate estimation of the degree of spatial

correlation. While such informed design is not always

possible, even cursory consideration of these ideas

should lead to improved sampling designs and thus

more accurate models.

ADVANTAGES OF BAYESIAN SPATIAL AND SPATIOTEMPORAL

HIERARCHICAL MODELS

Numerous examples demonstrate the advantages of

accounting for spatial and/or temporal correlation in

Bayesian hierarchical models for ecological problems. In

the area of species distribution, a series of papers by

Gelfand and coauthors (Gelfand et al. 2005, 2006,

Latimer et al. 2006) developed a complex hierarchical

modeling framework that led to new insights into the

spatial distributions of 23 species of a flowering plant

family in South Africa. These authors showed that

accounting for spatial correlation facilitated the assess-

ment of the factors that impact species distributions,

produced accurate maps of species occurrence, and

allowed for honest assessment of uncertainty. This work

is a particularly good example of the advantages of a

Bayesian analysis. The Bayesian paradigm allows for in

depth exploration of a virtually unlimited set of results

through careful thought and collaboration between

ecologists and statisticians. One warning, however, is

that the South African species distribution analyses were

complex and required many person-hours to produce

results. While this work is a terrific example of the

possibilities of a Bayesian analysis, it is also an example of

the complexities involved in doing such a careful analysis.

In the area of disease ecology, Waller et al. (2007)

considered the county-specific incidence of Lyme disease

in humans for the northeastern United States. They

examined a suite of models ranging from a standard

least-squares independent-errors regression model to a

hierarchical Bayesian model that accounted for spatial

correlation among counties. The inclusion of spatial and

temporal components in the model led to new insights

into the spread of Lyme disease over space and time and

produced maps showing disease trends over space and

time. The Bayesian model also allowed for natural

incorporation of missing data; the model provided

estimates for sites where the predictors were known

but the response (Lyme disease counts) was unknown.

This work led to new insights into the factors that might

contribute to the spread of Lyme disease.

April 2009 575HIERARCHICAL MODELS IN ECOLOGY

In the area of wildlife disease, Farnsworth et al. (2006)

used spatial models to link spatial patterns to the scales

over which generating processes operate. They devel-

oped a Bayesian hierarchical model to relate scales of

deer movement to observed patterns of chronic wasting

disease (CWD) in mule deer in Colorado. The Bayesian

hierarchical model allowed for investigation of the

effects of covariates observed at different scales;

covariates for individual deer (e.g., sex and age) and

covariates observed across the landscape (e.g., percent-

age of low-elevation grassland habitat) were included in

the model for the probability than an individual deer

was infected by CWD. The modeling framework also

facilitated a comparison of models for CWD across

different scales of deer movement via a model for the

unexplained variability in the probability that an

individual deer was infected by CWD. The model with

the strongest support suggested that unexplained vari-

ability has a small-scale component. This led the authors

to suggest that future investigations into the spread of

chronic wasting disease should focus on processes that

operate at smaller, local-contact scales. For example,

deer congregate in smaller areas during the winter and

disperse across the landscape during the summer; thus

CWD may be spread more easily during the winter

months.

Hooten and Wikle (2008) considered the spread of an

invasive species over time and space. They demonstrated

that many insights can be gained via a spatiotemporal

model that incorporates a reaction–diffusion component

to model the spread of the invasive Eurasian Collared-

Dove in the United States. This paper demonstrates

another strength of the Bayesian approach as it allows

for a natural incorporation of partial differential

equation models, long used in mathematics but typically

not parameterized to allow for process and data error.

The Hooten and Wikle model allows for uncertainty and

nonlinearity in the diffusion model (process error) as

well as an error term that allows for both observer error

and small-scale spatiotemporal variability (data error).

The analyses provided a series of maps estimating the

extent of the Eurasian Collared-Dove invasion over time

for the southeastern United States. The authors

concluded that there is remaining variability associated

with the rate of species invasion and not attributable to

human population. This remaining variability has an

estimated spatial range of 1/10 the size of the United

States. Such conclusions allow biologists to do a

targeted search for other factors that might contribute

to the spread of this invasive species.

CHALLENGES AND EDUCATION

All of the spatial and spatiotemporal modeling cited

in the previous section involved close collaboration

between ecologists and statisticians. While such collab-

orations advance the fields of statistics and ecology,

statisticians need to develop more approachable inter-

faces to allow scientists to apply complex Bayesian

hierarchical models. As the field of Bayesian hierarchical

modeling has matured, software packages such as

WinBUGS (available online)2 have made it possible for

non-experts to implement the Markov chain Monte

Carlo methods required for estimation and inference in

many Bayesian hierarchical models. However, much

more work needs to be done in this area, particularly for

the more complex models that account for spatial

and/or temporal correlation.

In addition, a push for more modern statistical

education in ecology and other sciences is needed. In

the meantime, where can an ecologist learn more about

statistical models to account for spatial and/or temporal

correlation? An introduction to these issues with an

ecological focus is given in the book by Clark (2007).

Waller and Gotway (2004) focus on spatial statistics in

public health. Books that require a higher-level under-

standing of statistics but that are still quite accessible

include Banerjee et al. (2004), which focuses on Bayesian

hierarchical models for spatial data, and Schabenberger

and Gotway (2005), which provides a broad overview to

statistical methods for spatial data. Both these books

include sections on spatiotemporal modeling. Other

overviews of spatiotemporal models with an ecological

focus include Wikle (1993) and Wikle et al. (1998).

ACKNOWLEDGMENTS

This work was supported by National Science Foundationgrant DEB-0717367.

LITERATURE CITED

Banerjee, S., B. P. Carlin, and A. E. Gelfand. 2004. Hierarchicalmodeling and analysis for spatial data. Chapman andHall/CRC Press, Boca Raton, Florida, USA.

Clark, J. S. 2007. Models for ecological data: an introduction.Princeton University Press, Princeton, New Jersey, USA.

Cressie, N. A. C., C. A. Calder, J. S. Clark, J. M. Ver Hoef, andC. K. Wikle. 2009. Accounting for uncertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19:553–570.

Farnsworth, M. L., J. A. Hoeting, N. T. Hobbs, and M. W.Miller. 2006. Linking chronic wasting disease to mule deermovement scales: a hierarchical Bayesian approach. Ecolog-ical Applications 16:1026–1036.

Gelfand, A. E., A. M. Schmidt, S. Wu, J. A. Silander, Jr., A.Latimer, and A. G. Rebelo. 2005. Modelling species diversitythrough species level hierarchical modeling. Journal of theRoyal Statistical Society, Series C (Applied Statistics) 54:1–20.

Gelfand, A. E., J. A. Silander, Jr., S. Wu, A. Latimer, P. O.Lewis, A. G. Rebelok, and M. Holder. 2006. Explainingspecies distribution patterns through hierarchical modeling.Bayesian Analysis 1:41–92.

Hoeting, J. A., R. A. Davis, A. A. Merton, and S. E.Thompson. 2006. Model selection for geostatistical models.Ecological Applications 16:87–98.

Hooten, M., and C. Wikle. 2008. A hierarchical Bayesian non-linear spatio-temporal model for the spread of invasivespecies with application to the Eurasian Collared-Dove.Environmental and Ecological Statistics 15:59–70.

Irvine, K, A. I. Gitelman, and J. A. Hoeting. 2007. Spatialdesigns and properties of spatial correlation: effects on

2 hhttp://www.mrc-bsu.cam.ac.uk/bugsi

FORUM576Ecological Applications

Vol. 19, No. 3

covariance estimation. Journal of Agricultural, Biologicaland Environmental Statistics 12(4)450–469.

Latimer, A. M., S. Wu, A. E. Gelfand, and J. A. Silander, Jr.2006. Building statistical models to analyze species distribu-tions. Ecological Applications 16:33–50.

Ritter, K., and M. Leecaster. 2007. Multi-lag cluster designs forestimating the semivariogram for sediments affected byeffluent discharges offshore in San Diego. Environmentaland Ecological Statistics 14:41–53.

Schabenberger, O., and C. A. Gotway. 2005. Statisticalmethods for spatial data analysis. Chapman and Hall/CRC,Boca Raton, Florida, USA.

Waller, L. A., B. J. Goodwin, M. L. Wilson, R. S. Ostfeld, S.Marshall, and E. B. Hayes. 2007. Spatio-temporal patterns incounty-level incidence and reporting of Lyme disease in the

northeastern United States, 1990–2000. Environmental andEcological Statistics 14:83–100.

Waller, L., and C. Gotway. 2004. Applied spatial statistics forpublic health data. Wiley, New York, New York, USA.

Wikle, C. K. 2003. Hierarchical Bayesian models for predictingthe spread of ecological processes. Ecology 84:1382–1394.

Wikle, C. K., L. M. Berliner, and N. Cressie. 1998. HierarchicalBayesian space–time models. Environmental and EcologicalStatistics 5:117–154.

Xia, G., M. L. Miranda, and A. E. Gelfand. 2006. Approx-imately optimal spatial design approaches for environmentalhealth data. Environmetrics 17:363–385.

Zimmerman, D. L. 2006. Optimal network design for spatialprediction, covariance parameter estimation and empiricalprediction. Environmetrics 17:635–652.

________________________________

Ecological Applications, 19(3), 2009, pp. 577–581� 2009 by the Ecological Society of America

Hierarchical Bayesian statistics: merging experimental and modelingapproaches in ecology

KIONA OGLE1

Departments of Botany and Statistics, University of Wyoming, Laramie, Wyoming 82071 USA

INTRODUCTION

This is an exciting time in ecological research because

modern data analytical methods are allowing us to

address new and difficult problems. As noted by Cressie

et al. (2009), hierarchical statistical modeling provides a

statistically rigorous framework for synthesizing eco-

logical information. For example, hierarchical Bayesian

methods offer quantitative tools for explicitly integrat-

ing experimental and modeling approaches to address

important ecological problems that have eluded ecolo-

gists due to limitations imposed by classical approaches.

Fruitful interactions between ecologists and statisticians

have spawned dialogue and specific examples demon-

strating the utility of such modeling approaches in

ecology (e.g., Wikle 2003, Clark and Gelfand 2006a, b,

Ogle and Barber 2008), and I applaud Cressie et al. for

introducing ecologists to some of the fundamental

statistical and probability concepts underlying hierarch-

ical statistical modeling. Readers are also referred to

Ogle and Barber (2008) for a more in-depth treatment of

the hierarchical modeling framework, fundamental

probability results, and examples that illustrate the

advantages of this approach in plant physiological and

ecosystem ecology.

Hierarchical statistical modeling approaches are

promising for addressing complex ecological problems,

and I imagine that in the next 10–20 years, these

approaches will be commonly employed in ecological

data analysis. However, application of these approaches

requires appropriate training, but training opportunities

in hierarchical modeling methods that integrate exper-

imental and/or observational data with models are

lacking (Hobbs and Hilborn 2006, Hobbs et al. 2006,

Little 2006). Thus, overview papers such as those by

Cressie et al. (2009) and Ogle and Barber (2008) are

expected to stimulate interests and motivate new

curriculums that deliver training in modern, model-

based approaches to data analysis. Cressie et al. discuss

several strengths and limitations of hierarchical statis-

tical modeling, but no single topic is treated in great

detail. Thus, I expand upon the importance of the

‘‘process model’’ because I see this as a key element of

hierarchical statistical modeling that facilitates explicit

integration of experiments (data) and ecological theory

(models).

EXPERIMENTAL VS. MODELING APPROACHES

Ecologists generally take one of two approaches to

tackling scientific problems: experimental vs. modeling

(Herendeen 1988, Grimm 1994). Perhaps the most

common approach is to couch a study within an

experimental or sampling design framework that ‘‘con-

trols’’ for sources of variability, followed up by stand-

ard, frequentist-based hypothesis testing (Cottingham et

Manuscript received 26 March 2008; accepted 16 June 2008.Corresponding Editor: N. T. Hobbs. For reprints of this Forum,see footnote 1, p. 551.

1 E-mail: [email protected]

April 2009 577HIERARCHICAL MODELS IN ECOLOGY

al. 2005, Hobbs and Hilborn 2006, Stephens et al. 2007).

Alternatively, one may employ mathematical or simu-

lation models that may have weak ties to experimental

data and that are often used for hypothesis generation,

parameter estimation, or prediction (Pielou 1981,

Rastetter et al. 2003). Rarely are these two approaches

combined in such a way that considers all relevant

information (e.g., data), preserves sources of variability

introduced by the experimental or sampling design,

formalizes existing knowledge about the ecological

system through quantitative models, and simultaneously

allows for hypothesis testing, hypothesis generation, and

parameter estimation.

Although the experimental approach can yield im-

portant insight into the behavior of ecological systems, it

is a ‘‘slow’’ learning process that forces the experimen-

talist to design a study to fit a relatively restrictive

analysis framework (Little 2006). For example, a

particular study may yield multiple types of data

representing different biological processes operating at

diverse temporal and spatial scales. As Ogle and Barber

(2008) note, however, the data are often treated in a

piece-wise fashion where different components of the

data are analyzed independent of each other despite the

fact that all/most data arise from interconnected

processes. Moreover, data analysis tends to proceed

via simple analysis of variance (ANOVA) or regression

methods (Cottingham et al. 2005, Hobbs and Hilborn

2006) that assume linearity and normality of responses

(e.g., data) and parameters (e.g., treatment effects,

regression coefficients). Depending on the context, more

sophisticated analyses in the form of hierarchical

statistical modeling may be applied. It should be noted

that most hierarchical modeling methods can be

implemented within different frameworks such as

Bayesian or maximum likelihood. Examples include

repeated-measurement or longitudinal data models

(Lindsey 1993, Diggle et al. 2002), multilevel models

(Goldstein 2002, Gelman and Hill 2006), or, more

generally, linear, nonlinear, or generalized linear (for

non-Gaussian data) mixed-effects models (e.g., Searle et

al. 1992, Davidian and Giltinan 1995, Gelman and Hill

2006). Such hierarchical modeling methods, however,

are often not employed by ecologists even though the

data structure may call for such methods (e.g., Potvin et

al. 1990, Peek et al. 2002).

Cressie et al. (2009) point out an important short-

coming of non-hierarchical, classical analyses such as

ANOVA and regression. That is, these methods assume

that ‘‘the uncertainty lies with the data and is due to

sampling and measurement.’’ The hierarchical versions

that include random effects, however, allow for addi-

tional sources of uncertainty due to, for example,

temporal, spatial, or individual variability that is

separate from measurement error. When considering

real ecological systems, especially in field settings, a large

fraction of the unexplained variability may be due to

process error. Process error exists for two primary and

interrelated reasons. First, other influential factors (e.g.,

covariates) that vary by, for example, individual (or

‘‘subject’’), time, or location, are often not measured or

not considered in the analysis. Second, it is impossible to

define a statistical or mathematical model that describes

the ecological system exactly. Importantly, classical data

analysis approaches may be inappropriate in many

settings because ecological systems give rise to multiple

sources of uncertainty and involve multiple, interacting,

nonlinear, and potentially non-Gaussian processes.

On the other end of the spectrum lies the modeling

approach. The application of process models in ecology

has a rich history stemming back to, for example,

theoretical models in population and community ecol-

ogy (see reviews by Getz 1998, Jackson et al. 2000) to

ecological systems modeling (Patten 1971, Jørgensen

1986) such as terrestrial ecosystem models (see reviews

by Ulanowicz 1988, Hurtt et al. 1998). Process models

can range from simple, empirical equations for recon-

structing observed patterns to detailed formulations

representing underlying mechanisms and interactions.

Thus, they can take on many different forms including,

for example, stationary and lumped parameter models

to deterministic or stochastic differential equation

models (e.g., Jørgensen 1986, Gertsev and Gertseva

2004). Such models are often applied for heuristic

purposes, and the process of going from a conceptual

model to a set of mathematical equations or computer

code forces one to quantify existing knowledge, thus

improving one’s understanding of the ecological system

(Grimm 1994, Rastetter et al. 2003).

Ecological process models have been used in different

capacities. Some are employed as tools for gaining

inductive insight about an ecological system based on

the mathematical structure and behavior of the model

(e.g., stability analysis; Getz 1998). When used in this

capacity, the models are generally evaluated indepen-

dent of actual data. Other applications use models to

make predictions about an ecological system, and the

predictions may be compared against data as a form of

model validation (Jørgensen 1986, Jackson et al. 2000).

Ogle and Barber (2008) note that the functional forms

and parameter values defining such models are often

derived from empirical information, but a model of even

moderate complexity is rarely rigorously fit to data. For

more detailed models, parameter values are commonly

obtained via ‘‘tuning’’ so that the model adequately fits a

particular data set or, more often, a summary of such

data (e.g., sample means). This approach is frequently

referred to as ‘‘forward modeling,’’ and comparatively

little emphasis is placed on rigorous statistical quanti-

fication of parameter and process uncertainty.

Thus, both ecological process modeling and hierarch-

ical statistical modeling are not new. The new and

exciting aspect of hierarchical statistical modeling—and

especially hierarchical Bayesian modeling as presented

by Cressie et al. (2009) and Ogle and Barber (2008)—is

the explicit integration of data obtained from exper-

FORUM578Ecological Applications

Vol. 19, No. 3

imental and/or observational studies with process

models that encapsulate our understanding of the

ecological system. This approach facilitates model-based

inference and permits us to break free from the relatively

restrictive framework of classical hypothesis testing and

design-based inference. In particular, it allows us to

acknowledge a richer understanding of the ecological

system(s), and the process model itself embodies a suite

of hypotheses about how the system behaves (e.g.,

Hobbs and Hilborn 2006). The hierarchical Bayesian

framework enables simultaneous analysis of diverse

sources of data within the context of an ecological

process model or models, thereby providing a very

flexible approach to data analysis. I see such flexibility as

a major strength of this approach because it can

accelerate our ability to gain new ecological insights

and develop and test new hypotheses.

INTEGRATION OF EXPERIMENTS AND MODELS

In the previous section, I refer to the ‘‘process model’’

as a mathematical or simulation model that can be

developed and applied independently of the hierarchical

statistical model. This terminology is different from that

in Cressie et al. (2009), and for clarity, I will refer to the

‘‘process model’’ as defined in their paper (i.e., [E jPE],

their Eqs. 1 and 2) as the ‘‘probabilitistic process

model.’’ The probabilitistic process model is a key

element of the hierarchical modeling framework because

it provides a direct link between the data and the

ecological process model. Cressie et al. use the harbor

seal haul-out example to nicely illustrate aspects of

hierarchical statistical modeling, but the process model

that they employ is highly empirical as it simply

describes the log of expected seal counts as a linear

function of three covariates and their squared terms. As

an ecologist who works with diverse and ‘‘messy’’ data

describing complex processes, I feel that their example

does not sufficiently illustrate the strengths of the

process model. This is what I emphasize here, and my

intention is to provide a conceptual overview that is

accessible to ecologists who may not have a strong

background in ecological modeling, statistical modeling,

or probability theory.

Let us begin by considering the right-hand side of Eqs.

2 and 4 in Cressie et al. (2009), where [D jE, PD] is the

data model, [E jPE] is the probabilistic process model,

and [PD, PE] is the parameter model (i.e., prior). For the

data model, one can think of an observed quantity (i.e.,

data) as ‘‘varying around’’ the true quantity (i.e., latent

process) plus observation or measurement error. For the

probabilistic process model, one can think of the true or

latent process—something that we can never observe

directly—as varying around an expected process plus

process error. For both models, the error terms are

described by probability distributions that quantify

variability in the measurement and process errors. Note

that the errors do not have to be additive; for example,

one could assume multiplicative errors, but a log

transformation would result in additive errors. Once

we have accounted for variability in the data explained

by the latent processes, then the remaining variability is

expected to reflect measurement errors. It is often

reasonable to assume that these errors are independent;

however, there are situations where knowledge about

the measurement process indicates that dependent errors

may be more appropriate (e.g., Ogle and Barber 2008).

Conversely, the process errors are likely to exhibit

greater ‘‘structure’’ such that they may be correlated in

time or space (e.g., Wikle et al. 1998, Wikle 2003); they

may also describe random effects reflecting uncertainty

introduced at various levels associated with data

collection (e.g., Clark et al. 2004). It is often the case

that these effects are nested (e.g., individual effects

nested within plot, plot effects nested within site, and

overall site effects). In most settings, structured process

errors are appropriate and follow from the experimental

or sampling design. Moreover, incorporation of struc-

ture is often necessary for separating measurement and

process errors, thereby avoiding identifiability problems.

Alternatively, random effects describing, for example,

individual or group effects can be directly incorporated

into the ecological process model (Ogle and Barber

2008), and process model parameters (PE) may be

modeled hierarchically. For example, Schneider (2006)

assumed that the parameters in a differential equation

model of plant growth differed between plots (q). They

observed several plots composed of single genotypes (g)

and modeled the plot-specific parameters (PEq) as

coming from a population described by genotype-

specific parameters (PEg(q)). The decision to incorporate

random effects into the processes errors vs. the process

parameters (or both) will depend on the problem at

hand and on one’s assumptions about the process

model.

Let us return to the probabilistic process model as

describing how the latent process(es) vary around an

‘‘expected process’’ plus process error. The expected

process is the ecological process model, and specification

of this component can be the most challenging part of

assembling a hierarchical Bayesian model. However, this

is where the vast knowledge accumulated by ‘‘ecological

modelers’’ can be incorporated. It is at this stage that

one formalizes their understanding of the ecological

system into a consistent set of mathematical equations,

yielding a model for the expected process. In construct-

ing this model, one is challenged with balancing detail

with simplicity such that important components, pro-

cesses, and interactions are incorporated in such a way

that pieces of the model can be directly related to

observable quantities. Simplicity is also key because it is

important that one understands the model’s behavior

with respect to how model outputs are affected by

functional forms (or specific equations) and parameter

values in these equations.

The goal is to arrive at a process model or set of

process models—representing alternative hypotheses

April 2009 579HIERARCHICAL MODELS IN ECOLOGY

about the system—that can be rigorously informed by

data. One may choose to define an empirical process

model that essentially describes a more traditional

ANOVA or linear regression model, but such a model

will provide limited insight into underlying mechanisms.

Conversely, I advocate using models that reflect

hypothesized nonlinearities and interactions and that

may represent processes operating at different bio-

logical, spatial, or temporal scales. The hierarchical

statistical modeling framework, and especially the

hierarchical Bayesian framework, can accommodate

this complexity by simultaneously considering data

collected at different scales that are directly related to

model inputs (e.g., parameters) and/or model outputs.

Of course, the scientific objectives of a particular study

should dictate the structure of the process model in the

same way that the objectives lead to a particular

experimental or sampling design, but neither should

determine the objectives, as is often the case in classical,

design-based inference.

In summary, the above discussion highlights the fact

that data obtained under an experimental design

framework can be analyzed in a hierarchical statistical

modeling framework that integrates all relevant data

with process models constructed to capture key eco-

logical interactions and nonlinearities. In particular, I

find that the hierarchical Bayesian approach provides

the most straightforward method for integrating diverse

data and process models. Although parameter estima-

tion is often the goal of hierarchical Bayesian modeling,

hypothesis testing can also proceed by evaluating

posterior distributions of parameters describing hy-

pothesized relationships or responses. More impor-

tantly, hypothesis testing and parameter estimation can

occur within the context of process models that embody

our understanding of the ecological system.

I consider the strengths of hierarchical statistical

modeling, and hierarchical Bayesian modeling in partic-

ular, to significantly outweigh the limitations because

these methods can bring together all sources of

information to bear on a particular problem, thereby

accelerating our ability to study and learn about real

ecological systems. The primary issues that the field of

ecology faces with respect to this exciting and powerful

approach to ecological data analysis is adequate training

and education. In the absence of sufficient training,

these methods may be underutilized, misunderstood, or

incorrectly applied.

ACKNOWLEDGMENTS

Much of my thought on hierarchical statistical modeling andBayesian methods has evolved through repeated discussionswith Jarrett Barber, and I thank Jarrett Barber and JessicaCable for valuable comments on this manuscript.

LITERATURE CITED

Clark, J. S., and A. E. Gelfand. 2006a. A future for models anddata in environmental science. Trends in Ecology andEvolution 21:375–380.

Clark, J. S., and A. E. Gelfand. 2006b. Hierarchical modellingfor the environmental sciences: statistical methods andapplications. Oxford University Press, Oxford, UK.

Clark, J. S., S. LaDeau, and I. Ibanez. 2004. Fecundity of treesand the colonization–competition hypothesis. EcologicalMonographs 74:415–442.

Cottingham, K. L., J. T. Lennon, and B. L. Brown. 2005.Knowing when to draw the line: designing more informativeecological experiments. Frontiers in Ecology and theEnvironment 3:145–152.

Cressie, N. A. C., C. A. Calder, J. S. Clark, J. M. Ver Hoef, andC. K. Wikle. 2009. Accounting for uncertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19:553–570.

Davidian, M., and D. M. Giltinan. 1995. Nonlinear models forrepeated measurement data. Chapman and Hall/CRC Press,Boca Raton, Florida, USA.

Diggle, P. J., P. Heagerty, K.-Y. Liange, and S. L. Zeger. 2002.Analysis of longitudinal data. Oxford University Press, NewYork, New York, USA.

Gelman, A., and J. Hill. 2006. Data analysis using regressionand multilevel/hierarchical models. Cambridge UniversityPress, New York, New York, USA.

Gertsev, V. I., and V. V. Gertseva. 2004. Classification ofmathematical models in ecology. Ecological Modelling 178:329–334.

Getz, W. M. 1998. An introspection on the art of modeling inpopulation ecology. BioScience 48:540–552.

Goldstein, H. 2002. Multilevel statistical models. HodderArnold, London, UK.

Grimm, V. 1994. Mathematical models and understanding inecology. Ecological Modelling 75:641–651.

Herendeen, R. A. 1988. Role of models in ecology. EcologicalModelling 43:133–136.

Hobbs, N. T., and R. Hilborn. 2006. Alternatives to statisticalhypothesis testing in ecology: a guide to self teaching.Ecological Applications 16:5–19.

Hobbs, N. T., S. Twombly, and D. S. Schimel. 2006. Deepeningecological insights using contemporary statistics. EcologicalApplications 16:3–4.

Hurtt, G. C., P. R. Moorcroft, S. W. Pacala, and S. A. Levin.1998. Terrestrial models and global change: challenges for thefuture. Global Change Biology 4:581–590.

Jackson, L. J., A. S. Trebitz, and K. L. Cottingham. 2000. Anintroduction to the practice of ecological modeling. Bio-Science 50:694–706.

Jørgensen, S. E. 1986. Fundamentals of ecological modelling.Elsevier, Amsterdam, The Netherlands.

Lindsey, J. K. 1993. Models for repeated measurements.Oxford University Press, New York, New York, USA.

Little, R. J. 2006. Calibrated Bayes: a Bayes/frequentistroadmap. American Statistician 60:213–223.

Ogle, K., and J. J. Barber. 2008. Bayesian data-modelintegration in plant physiological and ecosystem ecology.Progress in Botany 69:281–311.

Patten, B. C. 1971. Systems analysis and simulation in ecology.Volume I. Academic Press, New York, New York, USA.

Peek, M. S., E. Russek-Cohen, D. A. Wait, and I. N. Forseth.2002. Physiological response curve analysis using nonlinearmixed models. Oecologia 132:175–180.

Pielou, E. C. 1981. The usefulness of ecological models: a stock-taking. Quarterly Review of Biology 56:17–31.

Potvin, C., M. J. Lechowicz, and S. Tardif. 1990. The statisticalanalysis of ecophysiological response curves obtained fromexperiments involving repeated measures. Ecology 71:1389–1400.

Rastetter, E. B., J. D. Aber, D. P. C. Peters, D. S. Ojima, andI. C. Burke. 2003. Using mechanistic models to scaleecological processes across space and time. BioScience 53:68–76.

FORUM580Ecological Applications

Vol. 19, No. 3

Schneider, M. K., R. Law, and J. B. Illian. 2006. Quantificationof neighbourhood-dependent plant growth by Bayesianhierarchical modelling. Journal of Ecology 94:310–321.

Searle, S. R., G. Casella, and C. E. McCulloch. 1992. Variancecomponents. Wiley, New York, New York, USA.

Stephens, P. A., S. W. Buskirk, and C. Martınez del Rio. 2007.Inference in ecology and evolution. Trends in Ecology andEvolution 22:192–197.

Ulanowicz, R. E. 1988. On the importance of higher-level

models in ecology. Ecological Modelling 43:45–56.

Wikle, C. K. 2003. Hierarchical Bayesian models for predicting

the spread of ecological processes. Ecology 84:1382–1394.

Wikle, C. K., L. M. Berliner, and N. Cressie. 1998. Hierarchical

Bayesian space-time models. Environmental and Ecological

Statistics 5:117–154.

________________________________

Ecological Applications, 19(3), 2009, pp. 581–584� 2009 by the Ecological Society of America

Bayesian methods for hierarchical models:Are ecologists making a Faustian bargain?

SUBHASH R. LELE1,3

AND BRIAN DENNIS2

1Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, Alberta T6G2G1 Canada2Department of Fish and Wildlife Resources and Department of Statistics, University of Idaho, Moscow, Idaho 83844-1136 USA

It is unquestionably true that hierarchical models

represent an order of magnitude increase in the scope

and complexity of models for ecological data. The past

decade has seen a tremendous expansion of applications

of hierarchical models in ecology. The expansion was

primarily due to the advent of the Bayesian computa-

tional methods. We congratulate the authors for writing

a clear summary of hierarchical models in ecology.

While we agree that hierarchical models are highly

useful to ecology, we have reservations about the

Bayesian principles of statistical inference commonly

used in the analysis of these models. One of the major

reasons why scientists use Bayesian analysis for hier-

archical models is the myth that for all practical

purposes, the only feasible way to fit hierarchical models

is Bayesian. Cressie et al. (2009) do perfunctorily

mention the frequentist approaches but are quick to

launch into an extensive review of the Bayesian analyses.

Frequentist inferences for hierarchical models such as

those based on maximum likelihood are beginning to

catch up in ease and feasibility (De Valpine 2004, Lele et

al. 2007). The recent ‘‘data cloning’’ algorithm, for

instance, ‘‘tricks’’ the Bayesian MCMC setup into

providing maximum likelihood estimates and their

standard errors (Lele et al. 2007). A natural question

that a scientist should ask is: if one has a hierarchical

model for which full frequentist (say, based on ML

estimation) as well as Bayesian inferences are available,

which should be used and why? This can only be

answered based on the philosophical underpinnings. The

convenience criterion, commonly used to justify the use

of Bayesian approach, no longer applies, given the

recent advances in the frequentist computational ap-

proaches. Although the Bayesian computational algo-

rithms made statistical inference for such models

possible, it is not clear to us that the Bayesian inferential

philosophy necessarily leads to good science. In return

for seemingly confident inferences about important

quantities in the face of poor data and vast natural

uncertainties, are ecologists making a Faustian bargain?

We begin by stating our reservations about the

scientific philosophy advocated by the authors. The

authors claim that modeling is for synthesis of informa-

tion (Cressie 2009). Furthermore, they claim that

subjectivity is unavoidable. We completely disagree with

both these statements. In our opinion, the fundamental

goal of modeling in science is to understand the

mechanisms underlying the natural phenomena. Models

are quantitative hypotheses about mechanisms that help

us connect our prospective understandings to observable

phenomena. Certainly in the sociology of the conduct of

science, subjectivity often enters in the array of

mechanisms hypothesized. However, good scientists are

trained rigorously toward considering as many alter-

native explanations as imagination allows, with the

ultimate filters being consistency with observation,

experiment, and previous reliable knowledge. Introduc-

ing more subjectivity into the process of acquiring

reliable knowledge introduces confounding factors in

the empirical filtering of hypotheses, and so as scientists,

we should be striving to reduce subjectivity instead of

increasing it. Just because there is subjectivity in

hypothesizing mechanisms, it does not give us a free

Manuscript received 18 March 2008; revised 8 July 2008;accepted 24 July 2008. Corresponding Editor: N. T. Hobbs. Forreprints of this Forum, see footnote 1, p. 551.

3 E-mail: [email protected]

April 2009 581HIERARCHICAL MODELS IN ECOLOGY

pass to introduce subjectivity in testing those mecha-

nisms against the data.

Obtaining hard, highly informative data requires

substantial resources and time. Can expert opinion play

a role in inference? Eliciting prior distributions from

experts for use in Bayesian statistical analysis has often

been suggested for incorporating expert knowledge.

However, eliciting priors is more art than science. Aside

from this operational difficulty, a far more serious

problem is in deciding ‘‘who is an expert.’’ In the news

media, ‘‘experts’’ are available to offer sound bites

favoring any side of any issue. As scientists, we might

insist that expert opinion should be calibrated against

reality. Furthermore, the weight the expert opinion

receives in the statistical analysis should be based on the

amount and quality of information such an expert

brings to the table. Bayesian analysis lacks such explicit

quantification of the expertness.

We do believe that expert opinion can and should be

brought into ecological analyses (Lele and Das 2000,

Lele 2004), although not by Bayesian methods. Re-

cently, Lele and Allen (2006) showed how expert

opinion can be incorporated using a frequentist frame-

work by eliciting data instead of a prior. The method-

ology suggested in Lele and Allen (2006) automatically

weighs expert opinion and hard data according to their

Fisher information about the parameters of the process.

The expert who brings in only ‘‘noise’’ and no

information automatically gets zero weight in the

analysis. Provided the expert opinion is truly informa-

tive, the confidence intervals and prediction intervals

after incorporation of such opinion are shown to be

shorter than the ones that would be obtained without its

inclusion. It is thus possible to incorporate expert

opinion under the frequentist paradigm.

Hierarchical models are attractive for realistic model-

ing of complexity in nature. However, as a general

principle, the complexity of the model should be

matched by the information content of the data. As

the number of hierarchies in the model increases, the

ratio of information content to model complexity

necessarily decreases.

Unfortunately expositions of Bayesian methods for

hierarchical models have tended to emphasize practice

while deemphasizing the inferential principles involved.

Potential consequences of ignoring principles can be

severe when such data analyses are used in policy

making. In the following, we discuss some myths and

misconceptions about Bayesian inference.

‘‘Flat’’ or ‘‘objective’’ priors lead to desirable frequentist

properties.—Many applied statisticians and ecologists

believe that flat or non-informative priors produce

Bayesian credible intervals that have properties similar

to the frequentist confidence intervals. This idea has

been touted by the advocates of the Bayesian approach

as an important point of reassurance. An irony in this

claim is that Bayesian inference seems to be justified

under frequentist principles. However, the claim is plain

wrong. The Bayesian credible intervals obtained under

flat priors can have seriously incorrect frequentist

coverage properties; the actual coverage can be sub-

stantially smaller or larger than the nominal coverage

(Mitchell 1967, Heinrich 2005; D. A. S. Fraser, N. Reid,

E. Marras, and G. Y. Yi, unpublished manuscript). A

practicing scientist should ask: Under Bayesian infer-

ential principles, what statistical properties are consid-

ered desirable for credible intervals and how does one

assure them in practice?

Non-informative priors are unique.—Many ecologists

believe that the flat priors are the only kind of

distributions that are ‘‘objective’’ priors. In fact, in

Bayesian statistics the definition of non-informative

prior has been under debate since the 1930s with no

resolution. There are many different types of proper and

improper distributions that are considered ‘‘non-infor-

mative.’’ We highly recommend that ecologists read

Chapter 5 of Press (2003) and Chapter 6 of Barnett

(1999) for a non-mathematical, easily accessible dis-

cussion on the issue of non-informative priors. For a

quick summary, see Cox (2005). Furthermore, it is also

known that different ‘‘non-informative’’ priors lead to

different posterior distributions and hence different

scientific inferences (e.g., Tyul et al. 2008). The claim

that use of non-informative priors lets the data speak is

flatly incorrect. A scientist should ask: Which non-

informative priors should ecologists be using when

analyzing data with hierarchical models?

Credible intervals are more understandable than con-

fidence intervals.—The ‘‘objective’’ credible intervals

(i.e., formed with flat priors) do not have valid

frequentist interpretation in terms of coverage. For

subjective Bayesians, the interpretation of the coverage

is in terms of ‘‘personal belief probability’’. Neither of

these interpretations is valid for the ‘‘objective’’ Baye-

sian analysis. A scientist should ask: What is the correct

way to interpret ‘‘objective’’ credible intervals?

Bayesian prediction intervals are better than frequentist

prediction intervals.—Hierarchical models enable the

researcher to predict the unobserved states of the

system. The naıve frequentist prediction intervals, where

estimated parameter values are substituted as if they are

true parameter values, are correctly criticized for having

incorrect coverage probabilities. These intervals can be

corrected using bootstrap techniques (e.g., Laird and

Louis 1987). Such corrected intervals tend to have close

to nominal coverage. It is known that ‘‘objective’’

Bayesian credible intervals for parameters do not have

correct frequentist coverage. A scientist should ask: Are

prediction intervals obtained under the ‘‘objective’’

Bayesian approach guaranteed to have correct frequent-

ist properties? If not, how does one interpret the

‘‘probability’’ represented by these prediction intervals?

As sample size increases, the influence of prior

decreases rapidly.—For models that have multiple

parameters, it is usually difficult to specify non-

informative priors on all the parameters. The Bayesian

FORUM582Ecological Applications

Vol. 19, No. 3

scientists put non-informative priors on some of the

parameters and informative priors on the rest. Dennis

(2004) showed that the influence of an informative prior

can decrease extremely slowly as the sample size

increases when there are multiple parameters. We point

out that the hierarchical models being proposed in

ecology sometimes have enormous numbers of param-

eters. A scientist should ask: How does one evaluate the

influence of prior specification on the final inference?

The problem of identifiability can be surmounted by

specifying informative priors.—By definition, no exper-

imental data generated by the hierarchical model in

question can ever provide information about the non-

identifiable parameters. If this is the case, then how can

anyone have an ‘‘informed’’ guess about something that

can never be observed? Furthermore, as noted above

such ‘‘informative’’ priors can be inordinately influen-

tial, even for large samples. A scientist should ask: How

exactly does one choose a prior for a non-identifiable

parameter, and in what sense is specifying an informa-

tive prior a desirable solution for surmounting non-

identifiability?

Influence of priors on the final inference can be judged

from plots of the marginal posterior distributions.—The

true influence of the priors on the final inference is

manifested in how much the joint posterior distribution

differs from the joint prior. In practice, however,

influence of priors is judged by looking at the plots of

the marginal priors and posteriors. The marginalization

paradox (Dawid et al. 1973) suggests that the practice

could fail in spectacular ways. A scientist should ask:

How should one judge the influence of the specification

of the prior distribution on the final inference?

Bayesian posterior distributions can be used for

checking model adequacy.—Cressie et al. (2009) correctly

emphasize the importance of checking for model

adequacy in statistical inference. They suggest using

goodness of fit type tests on the Bayesian predictive

distribution. A scientist should ask: If such a test

suggests that the model is inadequate, how does one

know if the error is in the form of the likelihood function

or in the specification of the prior distribution?

Reporting sensitivity of the inferences to the specifica-

tion of the priors is adequate for scientific inference.—

Cressie et al. (2009), as well as many other Bayesian

analysts, suggest that one should conduct sensitivity of

the inferences to the specification of the priors. In our

opinion, it is not enough to simply report that the

inferences are sensitive. Inferences are guaranteed to be

sensitive to some priors and guaranteed to be non-

sensitive to some other priors. What is needed is a

suggestion as to what should be done if the inferences

are sensitive. A scientist should ask: If the inferences

prove sensitive to the particular choice of a prior, what

recourse does the researcher have?

Scientific method is better served by Bayesian ap-

proaches.—One of the most desirable properties of a

scientific study is that it be reproducible (Chang 2008).

The frequentist error rates, either the coverage proba-

bilities or probabilities of weak and misleading evidence

(Royall 2000), inform the scientists about the reprodu-

cibility of their results. A scientist should ask: How does

one quantify reproducibility when using ‘‘objective’’

Bayesian approach?

In summary, we applaud the authors for furthering

the discussion of use of hierarchical models in ecology.

While hierarchical models are useful, we cannot avoid

questioning the quality of inference that results from

hierarchy proliferation. Royall (2000) quantifies the

concept of weak inference in terms of probability of

weak evidence where, based on the observed data, one

cannot distinguish between different mechanisms with

enough confidence. We surmise that the probability of

weak evidence becomes unacceptably large as the

number of unobserved states increases. We feel that

scientists should be wary of introducing too many

hierarchies in the modeling framework. Furthermore,

we also have difficulties with the Bayesian approach

commonly used in the analysis of hierarchical models.

We suggest that the Bayesian approach is neither needed

nor is desirable to conduct proper statistical analysis of

hierarchical models. The alternative approaches based

on the frequentist philosophy of science should be

considered in analyzing the hierarchical models.

LITERATURE CITED

Barnett, V. 1999. Comparative statistical inference. John Wiley,New York, New York, USA.

Chang, K. 2008. Nobel winner retracts research paper. NewYork Times, 10 March 2008.

Cox, D. R. 2005. Frequentist and Bayesian statistics: a critique.In L. Louis and M. K. Unel, editors. Proceedings of theStatistical Problems in Particle Physics, Astrophysics andCosmology. Imperial College Press, UK. hhttp://www.physics.ox.ac.uk/phystat05/proceedings/default.htmi

Cressie, N. A. C., C. A. Calder, J. S. Clark, J. M. Ver Hoef, andC. K. Wikle. 2009. Accounting for uncertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19:553–570.

Dawid, A. P., M. Stone, and J. V. Zidek. 1973. Marginalizationparadoxes in Bayesian and structural inference. Journal ofthe Royal Statistical Society B 35:189–233.

Dennis, B. 2004. Statistics and the scientific method in ecology.Pages 327–378 in M. L. Taper and S. R. Lele, editors. Thenature of scientific evidence: statistical, empirical andphilosophical considerations. University of Chicago Press,Chicago, Illinois, USA.

De Valpine, P. 2004. Monte Carlo state-space likelihoods byweighted kernel density estimation. Journal of the AmericanStatistical Association 99:523–536.

Heinrich, J. 2005. The Bayesian approach to setting limits: whatto avoid. In L. Louis and M. K. Unel, editors. Proceedings ofthe Statistical Problems in Particle Physics, Astrophysics andCosmology. Imperial College Press, UK. hhttp://www.physics.ox.ac.uk/phystat05/proceedings/default.htmi

Laird, N. M., and T. A. Louis. 1987. Empirical Bayesconfidence intervals based on bootstrap samples. Journal ofthe American Statistical Association 82:739–750.

Lele, S. R. 2004. Elicit data, not prior: on using expert opinionin ecological studies. Pages 410–436 in M. L. Taper and S. R.Lele, editors. The nature of scientific evidence: statistical,empirical and philosophical considerations. University ofChicago Press, Chicago, Illinois, USA.

April 2009 583HIERARCHICAL MODELS IN ECOLOGY

Lele, S. R., and K. L. Allen. 2006. On using expert opinion inecological analyses: a frequentist approach. Environmetrics17:683–704.

Lele, S. R., and A. Das. 2000. Elicited data and incorporationof expert opinion for statistical inference in spatial studies.Mathematical Geology 32:465–487.

Lele, S. R., B. Dennis, and F. Lutscher. 2007. Data cloning:easy maximum likelihood estimation for complex ecologicalmodels using Bayesian Markov chain Monte Carlo methods.Ecology Letters 10:551–563.

Mitchell, A. F. S. 1967. Discussion of paper by I. J. Good.Journal of the Royal Statistical Society, Series B 29:423–424.

Press, S. J. 2003. Subjective and objective Bayesian statistics.Second edition. John Wiley, New York, New York, USA.

Royall, R. 2000. On the probability of observing misleadingstatistical evidence. Journal of the American StatisticalAssociation 95:760–768.

Tyul, F., R. Gerlach, and K. Mengersen. 2008 A comparison ofBayes–Laplace, Jeffreys, and other priors: the case of zeroevents. American Statistician 62:40–44.

________________________________

Ecological Applications, 19(3), 2009, pp. 584–588� 2009 by the Ecological Society of America

Shared challenges and common ground for Bayesian and classicalanalysis of hierarchical statistical models

PERRY DE VALPINE1

Department of Environmental Science, Policy and Management, 137 Mulford Hall #3114, University of California Berkeley,Berkeley, California 94720 USA

Let me begin by welcoming the paper by Cressie et al.

(2009) as an insightful overview of hierarchical statistical

modeling that will be valuable from both Bayesian and

classical perspectives. Statistical conclusions hinge on

the appropriateness of the mathematical models used to

represent hypotheses, and Cressie et al. explain many

merits of hierarchical models. My comments will high-

light the shared challenges and common ground of

Bayesian and classical analysis of hierarchical models;

give a likelihood theory complement to Cressie et al.’s

explanations of their merits; summarize methods for

maximum likelihood and ‘‘empirical Bayes’’ estimation

and incorporation of uncertainty; and explore some of

the claims of Bayesian advantages and classical limi-

tations.

Both Bayesian and classical analyses can and must

address the inherent challenges of inference and

prediction from noisy data, including comparing mod-

els, selecting a best model or combination of models,

deciding if a model is acceptable at all, avoiding over-

fitting and statistical ‘‘fishing’’ or ‘‘data dredging,’’ and

incorporating uncertainty. In practice, an appropriate

modeling framework trumps many other issues, includ-

ing choice of Bayesian or classical analysis.

The difference between Bayesian and classical philos-

ophies is that Bayesian analysis uses the mathematics of

probability distributions for model parameters, P as

defined by Cressie et al., while classical analysis does

not. It is generally agreed that Bayesian analysis must

define ‘‘probability of P’’ as ‘‘degree of belief for P’’ (or

‘‘subjective probability of P’’ or other synonymous

terms), whereas classical ‘‘probability’’ refers to a

frequency of outcomes over a long run (O’Hagan

1994). Bayesian analysis makes ‘‘probability’’ statements

about hypotheses, given data, that are formally weighted

degree-of-belief comparisons, while classical analysis

makes statements about frequencies with which different

hypotheses would have produced the data and, in some

approaches, hypothetical unobserved data. Classical

analysis encompasses much more than Neyman-Pearson

and/or Fisher hypothesis testing (which have been

critiqued for both their actual logical limitations when

correctly interpreted and their potential for misinter-

pretation; see Mayo and Spanos [2006] for extensions of

NP logic). For example, model comparison using

Akaike’s Information Criterion (AIC; Burnham and

Anderson 2002) takes a classical approach to parame-

ters. While Bayesian and classical approaches are

philosophically different, one can interpret results from

them in tandem (Efron 2005).

It is worth emphasizing that choosing a hierarchical

model is separate from choosing a Bayesian or classical

approach for parameters. In a hierarchical model, one

has parameters (P, either Bayesian or classical) that

define the distribution of unknown ecological states (E )

that define the distribution of data (D). I will call E the

latent variables and/or random effects, which are among

the many names for random variables whose values are

not directly known but in theory define relationships

among data values, to emphasize that hierarchical

models are mixed models (McCulloch and Searle

Manuscript received 20 March 2008; revised 8 July 2008;accepted 24 July 2008. Corresponding Editor: N. T. Hobbs. Forreprints of this Forum, see footnote 1, p. 551.

1 E-mail: [email protected]

FORUM584Ecological Applications

Vol. 19, No. 3

2001); as Cressie et al. point out, their example uses a

generalized linear mixed model. Having at least one E

level make a model hierarchical. Treating the parame-

ters, P, with degree-of-belief probability makes an

analysis Bayesian. Cressie et al. illustrate a general route

to a Bayesian analysis. A general route to a classical

analysis would typically include maximum likelihood

estimation of P under various models, with comparisons

and estimates of uncertainty made by likelihood ratios

(Royall 1997), likelihood ratio hypothesis tests andconfidence regions, AIC, parametric or nonparametric

bootstrapping, or other approaches.

Why are hierarchical models such a good idea for

either Bayesian or classical analysis? Cressie et al.explain the sensibility of conditionally nested hierarchies

of probability models in terms of ecological and

sampling processes. A complementary view is that the

likelihood of the parameters, P ¼ (PD, PE), given D, is

the integral of Cressie et al.’s Eq. 1 with respect to E:

LðPD;PEÞ ¼ ½DjPD;PE� ¼R½DjE;PD�½EjPE�dE: ð1Þ

This states that the probability of D given P is the sum

over all possible E values of the probability of (i) those Evalues given PE and (ii) D given PD and those E values.

This likelihood is the ‘‘[data j parameters]’’ mentioned by

Cressie et al. (2009). The asymptotic (as the amount of

data increases) behavior of this likelihood guarantees

convergence to the best P, a Gaussian shape for L, and

other properties (McCulloch and Searle 2001). By ‘‘best’’

P, one can mean either the ‘‘true’’ P or the P that

minimizes the theoretical Kullback-Leibler discrepancy

(Burnham and Anderson 2002). Likelihood asymptotics

give a theoretical bedrock for both Bayesian and

classical analysis. Thus, a more formal appeal of

hierarchical models is that they often define appropriate

likelihoods for P, which drive the soundness of results

from either Bayesian or classical analysis and oftenmake the two approaches yield similar conclusions.

What do I mean by saying that a good model trumps

many other considerations? Given a choice between

Bayesian analysis of a set of appropriate models andclassical analysis with a set of plainly inappropriate

models, or vice versa, I’ll usually take the one with the

appropriate models and then judge the pros and cons of

the analysis. Recently, such a choice has led to a

common pragmatic appeal of Bayesian analysis: for

many people it is currently the easiest framework for

analyzing a hierarchical model. With time, I suspect that

advances in software and algorithms will move us closer

to being able to match any model structure to any

analysis approach. Then one will be able to choose a set

of hierarchical models based on the ecology and conduct

either a Bayesian or classical analysis based on scientific

goals.

How would one obtain maximum likelihood results

for a hierarchical model such as Cressie et al.’s seal

example, so one can use likelihood ratios or AIC, for

example? For relatively simple models, generalized

linear mixed model software can accomplish this. For

more general models, one can use numerical integration

of (1) via either grid-based (Efron 1996) or Monte Carlo

integration approaches, reviewed by de Valpine (2004).

The latter include Monte Carlo expectation maximiza-

tion, Monte Carlo Newton-Raphson iterations, ‘‘direct’’

Monte Carlo integration with importance sampling,

sequential Monte Carlo integration (‘‘particle filters’’),

iterated Monte Carlo likelihood ratio approximations,

and Monte Carlo kernel likelihood approximations (de

Valpine 2004). More recent approaches include iterated

particle filtering (Ionides et al. 2006) and data cloning

(Lele et al. 2007). Each approach has pros and cons, just

as Monte Carlo simulation of Bayesian posteriors can be

done with relatively efficient or inefficient flavors of

Markov chain Monte Carlo (MCMC) algorithms, with

the best choice depending on the specific problem.

Monte Carlo kernel likelihood (MCKL) uses the full

Bayesian posterior, an appealing feature for those who

want to compare Bayesian and maximum likelihood

results.

Most of these algorithms use calculations that omit a

part of the likelihood known as the ‘‘normalizing

constant,’’ which must be estimated as a separate step.

This problem is mathematically identical to estimating

the marginal likelihood used in Bayes Factors after

simulating an MCMC posterior sample (de Valpine

2008). In summary, maximum likelihood and normaliz-

ing constant algorithms are highly feasible for many

hierarchical models, but current software makes Baye-

sian analysis more easily available for some models.

To a newcomer, the terminology that ‘‘empirical

Bayes’’ is a ‘‘non-Bayesian’’ analysis may seem baffling.

Typically, the empirical Bayes parameters would be the

maximum likelihood estimates, sometimes called the

‘‘plug-in’’ parameters (Cressie et al. 2009). They do not

require degree-of-belief probabilities and so are not

Bayesian. The confusing distinction is illustrated by

contrasting Efron (1986), who discussed empirical Bayes

as a frequentist method, and Little (2006), who called it

Bayesian in a ‘‘broad view’’ that encompasses ‘‘a large

class of practical frequentist methods with a Bayesian

interpretation.’’

Another potential confusion is that some purported

contrasts of Bayesian and classical results are con-

founded with contrasts of hierarchical and nonhierarch-

ical models, respectively. Empirical Bayes has its roots in

the estimation of E (not P): if a hierarchical model is

appropriate, then using the conditional distribution of E

given D for maximum likelihood P (empirical Bayes) or

a posterior for P (Bayes) can be better than using a

nonhierarchical maximum likelihood approach for E for

each unit of D (Morris 1983). Some contrasts use a

nonhierarchical model for the classical side and a

hierarchical model for the Bayesian (or empirical Bayes)

side. This issue is reflected in Cressie et al.’s (2009)

reference to a nonhierarchical analysis as the ‘‘usual

April 2009 585HIERARCHICAL MODELS IN ECOLOGY

likelihood analysis’’ and in examples in Carlin and Louis

(2000) and elsewhere.

How can we compare Bayesian and classical ap-

proaches? Some have proposed frequentist (classical)

evaluation of Bayesian analysis (Morris 1983, Rubin

1984, Robins et al. 2000). That is, the performance of

Bayesian analysis that treats parameters (P) with

subjective probability can be evaluated based on the

frequencies of outcomes over a ‘‘true’’ distribution of

hypothetical data, such as coverage of a ‘‘true’’ P or E

value by a credible region, which is naturally related to

P-value concepts (not to be confused here with model

parameters P). The motivation from a Bayesian view is

that ‘‘frequency calculations are useful for making

Bayesian statements scientific, . . . capable of being

shown wrong by empirical test’’ (Rubin 1984). Cressie

et al. (2009) appeal to this frequentist justification of

‘‘accurate’’ credible intervals, and such accuracy is

driven by the likelihood. The common ground that all

models should be scientifically rejectable using the same

performance currencies seems valuable. In contrast, a

‘‘pure’’ Bayesian would eschew any model testing based

on frequencies of hypothetical, unobserved data.

How do Bayesian and classical analyses of hierarchical

models each ‘‘incorporate uncertainty,’’ and how justi-

fied are the statements of Cressie et al. (2009) that the

Bayesian approach ‘‘captures the variability in the

parameters [P]’’ while plug-in estimates ‘‘do not account

properly for the variability,’’ and that frequentist

incorporation of uncertainty is limited in complexity?

Frequentist thinking has long recognized that estimated

parameters by definition give an optimistic picture of

model fit, and one must account for this in inference and

prediction. Cressie et al. mention quadratic likelihood

approximations, bootstrapping and cross-validation.

One could extend this list by including profile like-

lihoods, AIC, bootstrapping in empirical Bayes (Efron

1996), generalized cross-validation, generalized degrees

of freedom to account for over-fitting due to model

selection (Ye 1998), general covariance penalties (Efron

2004), and more. Nevertheless, in many situations a

Bayesian posterior sample will be the easiest picture of

uncertainty in P one can quickly generate, and indeed the

only feasible picture for relatively complex models. Since

neither approach is inherently superior at incorporating

uncertainty, and since this topic touches the core of much

statistical research, the pragmatic boundaries between

approaches are likely to change with time.

Although a Bayesian picture of uncertainty is valuable

and practical, it is not necessarily better for all purposes

than even simply ‘‘plugging-in’’ the maximum likelihood

estimate for P. For example, it turns out that using

posterior predictive intervals to test overall model fit can

be too conservative in the sense that they are guided by

the same data used to evaluate the model, so the model

is not rejected as often, in a frequency sense, as intended

by the analyst (Bayarri 2003, Gelfand 2003). Robins et

al. (2000) compared seven classical and ‘‘Bayesian’’ P

value approaches for model assessment. Maximum

likelihood (‘‘plug-in’’) P values were more accurate than

posterior predictive P values. The best approaches were

newer ones developed by Bayarri and Berger (2000),

who also concluded that the P value from maximum

likelihood estimates ‘‘seems superior’’ to the posterior

predictive P value, ‘‘which would seem to contradict the

common Bayesian intuition that it is better to account

for parameter uncertainty by using a posterior than by’’

plugging in the maximum likelihood estimate.

The related topics of model selection and model

validation or criticism are more important than ever

with the Pandora’s box of computational model-fitting

opened: we can fit a huge range of models to data—a

good problem to have—but must avoid being misled by

them. O’Hagan (2003) reviewed Bayesian ‘‘model

criticism’’ starting from a ‘‘growing unease that the

power of HSSS [hierarchical] modeling . . . was tempting

us to build models that we did not know how to criticize

effectively.’’ To give ecologists context about using DIC

for model selection (Cressie et al. 2009): Spiegelhalter et

al. (2002) presented DIC ‘‘tentatively’’ on ‘‘somewhat

heuristic’’ grounds that were motivated by classical

theory. It is also motivated pragmatically by using only

information available from MCMC output. AIC is

defined for a hierarchical model in terms of the

likelihood (1) and parameters, P, and BIC is an

asymptotic estimate of a Bayesian marginal likelihood,

but neither is available simply from MCMC output (de

Valpine 2008). The discussants of Spiegelhalter et al.

(2002) largely praised the DIC but raised many cautions

and questions about its performance. These fields are

sure to see both Bayesian and classical advances.

What should ecologists make of Cressie et al.’s claims

that Bayesian methods incorporate information ‘‘in a

coherent fashion’’ and give a ‘‘conceptually holistic

approach to inference’’? Based on other Bayesian usages

of incorporating information ‘‘coherently’’ (Carlin and

Louis 2000, Efron 2005), the first claim seems to refer

largely to formulating a likelihood function that includes

data related by random effects or from multiple sources

(studies or sampling procedures), so it is really a feature

of hierarchical models more than of Bayesian analysis. I

interpret at least an aspect of the second appeal to refer

to the handy practical situation that once you have a

posterior sample of all P and E dimensions at your

fingertips, much of the rest of your analysis involves

summarizing it in various ways without worrying about

breaking Bayesian rules. I think this is in part a valuable,

practical view, and in part overoptimistic. I have already

given the examples that calculating marginal likelihoods

for model comparisons requires additional computa-

tions; that many conceptually meaningful P values for

model criticism can be defined; that model selection will

see further development; and that the easily generated

posterior predictive intervals may represent over-fits to

the data. As an example of a more advanced approach,

Cressie et al. (2009) mention cross-validation, a form of

FORUM586Ecological Applications

Vol. 19, No. 3

which was recently used to criticize a Bayesian popula-

tion model for conservation (Snover 2008).

The above variety of considerations illustrate that a

thorough Bayesian analysis, beyond just a posterior, can

lead to a fairly complicated set of results requiring

careful judgment. This seems not very different from the

situation in classical analysis that structured probability

models define likelihoods, which provide a theoretical

core connecting related analyses. An example of how the

beguiling unity of treating both random effects and

parameters as Bayesian ‘‘parameters’’ led to mistakes in

justifying and using state–space fisheries models is

explained by de Valpine and Hilborn (2005).

Cressie et al. (2009) give an example of the practical

appeal of posterior probability summaries, such as 95%

credible intervals, for an E value. Credible intervals are

informative, but it is useful to note that such claims

about their interpretation revolve around the contrast

between degree-of-belief and frequency ‘‘probability.’’

I will touch briefly on some other aspects of

hierarchical model analysis set up by Cressie et al. While

the challenge of generating sound MCMC results varies

greatly between problems, I do not see computational

convergence as a major concern for the overall ap-

proaches. For full Bayesian analysis, a practical way to

assess the computational error for posterior summaries is

the Moving Block Bootstrap (Mignani and Rosa 2001).

The subjectivity of Bayesian priors seems to me a more

serious issue, but not for the most commonly mentioned

reasons. A more subtle problem than sensitivity to prior

parameters is that there is no such thing as a universally

flat prior, because flatness depends on the parameter-

ization of the model. Flat priors for r2, r, or 1/r2 will all

give different results, as will flat priors for k (population

growth rate) or log(k) in a population dynamics model.

In many cases, the difference in results may be small, but

nevertheless considering this issue should be a standard

step in applications. Finally, while I appreciate the

contrast between ‘‘curve fitting’’ and ‘‘formal statistical

modeling’’ (Cressie et al. 2009) as a gentle warm-up to the

rationale for hierarchical models, the distinction seems

limited. Even in a hierarchical model, one is estimating,

or ‘‘fitting,’’ parameters of ‘‘curves’’ as reflected by the

‘‘smooth curve’’ language of Cressie et al.’s seal example;

the advantage is doing so with a better model.

In summary, hierarchical models are an excellent

framework for analyzing data, and Cressie et al. should

go a long way towards helping ecologists adopt them.

Learning to formulate and interpret hierarchical models

involves learning to think clearly about stochastic

relationships among variables in complex systems.

Likelihood theory represents a major common ground

for most Bayesian and classical analysis methods and is

the reason they often give practically similar insights.

Both classical and Bayesian approaches have pros and

cons for pragmatism, performance, and interpretation.

My comments are not intended to tally points in a

debate, but rather to emphasize that inference and

prediction from noisy data are very hard problems.

Bayesian parameter distributions can give useful ac-

countings of uncertainty in many contexts. However,

some claims of Bayesian advantages based partly on

appeals to intuition do not hold up to theoretical

analysis, even if the intuitions have real merits.

While Bayesian approaches are practical for current

use, many methods exist for maximum likelihood

estimation and related analyses, including incorporation

of parameter uncertainty, that would work well for

many ecological hierarchical models. Gradually these

approaches will allow choice of analysis philosophy to

be based on scientific needs rather than having it tied by

pragmatism to choice of a hierarchical model structure.

Bayesian analysis will still be chosen for some important

problems. It will be important for more ecologists to

understand relationships between analysis methods to

reach shared understanding of statistical results, and

therefore of the evidence for hypotheses and support for

predictions from data.

ACKNOWLEDGMENTS

I thank Kevin Gross and Ray Hilborn for helpful commentson a draft. The perspectives and remaining flaws are my own.

LITERATURE CITED

Bayarri, M. J. 2003. Which ‘‘base’’ distribution for modelcriticism? Pages 445–448 in P. J. Green, N. L. Hjort, and S.Richardson, editors. Highly structured stochastic systems.Oxford University Press, Oxford, UK.

Bayarri, M. J., and J. O. Berger. 2000. P values for compositenull models. Journal of the American Statistical Association95:1127–1142.

Burnham, K. P., and D. R. Anderson. 2002. Model selectionand multimodel inference: a practical information-theoreticapproach. Springer, New York, New York, USA.

Carlin, B. P., and T. A. Louis. 2000. Bayes and empirical Bayesmethods for data analysis. Chapman and Hall/CRC, BocaRaton, Florida, USA.

Cressie, N. A. C., C. A. Calder, J. S. Clark, J. M. Ver Hoef, andC. K. Wikle. 2009. Accounting for uncertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19:553–570.

de Valpine, P. 2004. Monte Carlo state-space likelihoods byweighted posterior kernel density estimation. Journal of theAmerican Statistical Association 99:523–536.

de Valpine, P. 2008. Improved estimation of normalizingconstants from Markov chain Monte Carlo output. Journalof Computational and Graphical Statistics 17:333–351.

de Valpine, P., and R. Hilborn. 2005. State-space likelihoodsfor nonlinear fisheries time-series. Canadian Journal ofFisheries and Aquatic Sciences 62:1937–1952.

Efron, B. 1986. Why isn’t everyone a Bayesian? AmericanStatistician 40:1–5.

Efron, B. 1996. Empirical Bayes methods for combininglikelihoods. Journal of the American Statistical Association91:538–550.

Efron, B. 2004. The estimation of prediction error: covariancepenalties and cross-validation. Journal of the AmericanStatistical Association 99:619–632.

Efron, B. 2005. Bayesians, frequentists, and scientists. Journalof the American Statistical Association 100:1–5.

Gelfand, A. E. 2003. Some comments on model criticism. Pages449–453 in P. J. Green, N. L. Hjort, and S. Richardson,editors. Highly structured stochastic systems. Oxford Uni-versity Press, Oxford, UK.

April 2009 587HIERARCHICAL MODELS IN ECOLOGY

Ionides, E. L., C. Breto, and A. A. King. 2006. Inference fornonlinear dynamical systems. Proceedings of the NationalAcademy of Sciences (USA) 103:18438–18443.

Lele, S. R., B. Dennis, and F. Lutscher. 2007. Data cloning:easy maximum likelihood estimation for complex ecologicalmodels using Bayesian Markov chain Monte Carlo methods.Ecology Letters 10:551–563.

Little, R. J. 2006. Calibrated Bayes: a Bayes/frequentistroadmap. American Statistician 60:213–223.

Mayo, D. G., and A. Spanos. 2006. Severe testing as a basicconcept in a Neyman-Pearson philosophy of induction.British Journal for the Philosophy of Science 57:323–357.

McCulloch, C. E., and S. R. Searle. 2001. Generalized, linear,and mixed models. John Wiley and Sons, New York, NewYork, USA.

Mignani, S., and R. Rosa. 2001. Markov chain Monte Carlo instatistical mechanics: the problem of accuracy. Technomet-rics 43:347–355.

Morris, C. N. 1983. Parametric Empirical Bayes inference:theory and applications. Journal of the American StatisticalAssociation 78:47–55.

O’Hagan, A. 1994. Kendall’s advanced theory of statistics.Volume 2B: Bayesian inference. Oxford University Press,New York, New York, USA.

O’Hagan, A. 2003. HSSS model cricitism. Pages 423–444 inP. J. Green, N. L. Hjort, and S. Richardson, editors. Highlystructured stochastic systems. Oxford University Press, NewYork, New York, USA.

Robins, J. M., A. van der Vaart, and V. Ventura. 2000.Asymptotic distribution of P values in composite nullmodels. Journal of the American Statistical Association 95:1143–1156.

Royall, R. M. 1997. Statistical evidence: a likelihood paradigm.Chapman and Hall, New York, New York, USA.

Rubin, D. B. 1984. Bayesianly justifiable and relevant frequencycalculations for the applied statistician. Annals of Statistics12:1151–1172.

Snover, M. L. 2008. Comments on ‘‘Using Bayesian state-spacemodelling to assess the recovery and harvest potential of theHawaiian green sea turtle stock.’’ Ecological Modelling 212:545–549.

Spiegelhalter, D. J., N. G. Best, B. R. Carlin, and A. van derLinde. 2002. Bayesian measures of model complexity and fit.Journal of the Royal Statistical Society Series B—StatisticalMethodology 64:583–616.

Ye, J. M. 1998. On measuring and correcting the effects of datamining and model selection. Journal of the AmericanStatistical Association 93:120–131.

________________________________

Ecological Applications, 19(3), 2009, pp. 588–592� 2009 by the Ecological Society of America

Learning hierarchical models: advice for the rest of us

BEN BOLKER1

Zoology Department, University of Florida, Gainesville, Florida 32611-8525 USA

Cressie et al. (2009) propose that hierarchical model-

ing can revolutionize ecology, removing constraints that

have forced ecologists to oversimplify statistical models

and to ignore important distinctions between measure-

ment error, process error, and model uncertainty.

Hierarchical models are enormously flexible, allowing

ecologists to think about what questions they would like

to answer rather than what models their software can fit.

However, for ecologists to benefit from the advantages

of the hierarchical approach, they will need to give up

the safety of procedure-based statistics for a more

technically challenging, flexible, and subjective approach

to modeling.

Cressie et al. gloss over the huge challenge that this

transition poses for most ecologists. Most of my response

will address the practical issues that typical ecologists

(i.e., those without advanced course work or degrees in

statistics) will have to confront if they want to use

hierarchical models to understand their systems better.

BUGS AS A MODELING LANGUAGE

Cressie et al. (2009) briefly describe the various

implementations of the BUGS language (WinBUGS,OpenBUGS, JAGS, and so forth) that are available for

fitting hierarchical models. Customized code written in

lower-level languages is more flexible, and usually muchfaster, than these tools, but Cressie note that it is

‘‘considerably more tedious to implement’’ (i.e., prob-ably impossible for most ecologists). Ecologists who

want to use hierarchical models will usually need to starteither by finding existing software designed for a specific

class of problems (e.g., Okuyama and Bolker 2005) or bylearning to use some dialect of BUGS (Woodworth

2004, McCarthy 2007). While some researchers feel that

BUGS is too dangerous for inexperienced users whohave ‘‘limited understanding of the underlying assump-

tions’’ (Clark 2007:456), it is a very good way to getstarted using Markov chain Monte Carlo (MCMC)

methods for simple models. In particular, while Mc-Carthy (2007) touches only lightly on hierarchical

models, he provides a gentle introduction to Bayesian

statistics and MCMC, after which one can tackle Clark(2007) or Clark et al. (2006). Gelman and Hill (2006)

also present an enormous amount of useful knowledge

Manuscript received 7 April 2008; revised 12 June 2008;accepted 30 June 2008. Corresponding Editor: N. T. Hobbs. Forreprints of this Forum, see footnote 1, p. 551.

1 E-mail: [email protected]

FORUM588Ecological Applications

Vol. 19, No. 3

about hierarchical models, although they focus on

problems in social science. Automatic samplers like

WinBUGS may be very slow for larger models; after

getting the hang of basic models, ecologists should

probably follow Clark’s advice and graduate to coding

their own samplers for more complex problems.

An enormous advantage of the BUGS language is

that it separates a statistical model’s definition from its

implementation. Such ecological realities as nonnormal

distributions, nonlinear responses to predictor variables,

and multiple levels of error can easily be incorporated in

such models, using a straightforward descriptive lan-

guage. For example, a statistician would describe Ver

Hoef and Frost’s (2003) example, Poisson-distributed

variation in seal counts with random lognormal

variation across years, as

ht ; Normalðl;r2t Þ;

Yit ; Poisson½expðhtÞ�:ð1Þ

The first line says that ht, the (natural) logarithm of the

mean abundance in year t, is drawn from (;) a normal

distribution with mean l and variance r2; the second

line says that the observed abundance in site i in year t

(Yit) is drawn from a Poisson distribution with mean eht .

This notation gives us a compact and precise way to

describe statistical models, one that is more broadly

useful than the decomposition into normally distributed

variance components (e.g., Yij ¼ l þ ei þ ej) that

ecologists know from ANOVA designs. The BUGS

language closely parallels this notation: the statements

theta½t�;dnormðmu;tau½t�Þexptheta½t� ,�expðtheta½t�Þ

Y ½ i;t�;dpoisðexptheta½t�Þð2Þ

show the core of the BUGS code for the Ver Hoef and

Frost model. The only differences besides basic typog-

raphy (; instead of ;, ,– for assigning a value to a

variable) are that (1) BUGS describes the normal

distribution in terms of its inverse variance or precision,

s ¼ 1/r2, rather than the more familiar description in

terms of the variance, and (2) in some dialects of BUGS,

the value eht has to be computed in a separate statement.

These notations also solve another challenge raised by

hierarchical models. With increasing complexity of

statistical models comes decreased replicability. In the

old days of simple t tests, ANOVAs, and linear

regressions, it was usually easy to understand exactly

how authors had analyzed their data. Now, even though

raw data are much more likely to be available in

electronic form, replicating analyses can be much more

difficult. Precise statistical notation, and the lingua

franca of the BUGS language, offer a partial solution.

The same model run in any dialect of BUGS should give

qualitatively similar results. Although the stochastic

component of MCMC makes it impossible to replicate

results exactly across dialects, within a dialect one can

(and should) specify the starting seed for the random

number generator to ensure perfect replicability. Re-

viewers and editors should continue to hold authors to a

high standard of statistical replicability, even (especially)

when they present complex hierarchical models.

The recent development of new BUGS dialects like

JAGS has opened up the language, allowing cross-

validation of the software and driving innovation (much

as the development of the open source R language has

enhanced the older S-PLUS dialect of the S language).

Computational statisticians are also beginning to devel-

op intermediate-level tools (such as the MCMCpack

(Martin et al. 2008) and Umacs (Kerman 2007) packages

in R, or the HBC tool kit (available online),2 that will

help bridge the gap between the black box of BUGS and

the difficulty of coding models from scratch. With these

new software tools and the increasing availability of

high-performance computer clusters that can run models

over days or weeks if necessary, the computational

challenges of hierarchical models should diminish

greatly over the next few years.

CHALLENGES

Hierarchical models pose a set of challenges distinct

from both the research challenges described by Cressie et

al. and the computational difficulties I have just

described. Hierarchical models and other more general

frameworks such as maximum likelihood estimation

force researchers to worry about technical statistical

details such as whether they can really estimate the

parameters of a given model reliably, and how to make

inferences from the results. Hierarchical models are

much more finicky than classical procedures: ecologists

who want to use them will have consider technical

details such as choosing starting values, tuning opti-

mization procedures, and figuring out whether a

Markov chain sampler has converged.

Some of Cressie et al.’s challenges, such as determin-

ing the adequacy of MCMC sampling or assessing

convergence, are technical details that can be overcome

by increased computer power or better rules of thumb.

The much-criticized subjectivity of Bayesian statistics is

also, in my opinion, a relatively minor problem.

Ecologists are extreme pragmatists, and will tolerate

philosophical inconveniences, such as Bayesian subjec-

tivity, or the inferential contortions raised by the use of

null hypothesis testing. Reaching reasonably objective

conclusions with Bayesian methods is not trivial, but it is

no harder than many other technical challenges that

ecologists have mastered (Edwards 1996, Berger 2006).

The greatest challenge of hierarchical models is

pedagogical: how can ecologists learn to use the power

of hierarchical models, and especially of WinBUGS and

related tools, without getting themselves in trouble? The

WinBUGS manual famously starts by warning the user

to ‘‘Beware: MCMC sampling can be dangerous!’’ Even

2 hhttp://www.cs.utah.edu/;hal/HBC/i

April 2009 589HIERARCHICAL MODELS IN ECOLOGY

worse, the previous line of the manual states, ‘‘If there is

a problem, WinBUGS might just crash, which is not

very good, but it might well carry on and produce

answers that are wrong, which is even worse.’’ Clark

(2007) says that ‘‘turning nonstatisticians loose on

BUGs is like giving guns to children.’’ These warnings

are like Homeland Security threat advisories: they warn

of danger, but provide little guidance.

So how can ecologists avoid trouble? Some recom-

mendations:

1) Use common sense. Remember that there are no

free lunches: small, noisy data sets (such as many

graduate students collect) can only answer simple, well-

posed questions. The social and environmental scientists

who have driven the development of hierarchical models

typically have noisy but large data sets, where one has

thousands of data points from which to estimate

parameters. (One may be able to use noisy data where

there are a small number of samples in any particular

sampling unit, but in this case the design must include a

large number of sampling units.)

2) Consider identifiability. Cressie et al. point out that

incorporating redundant (unidentifiable) parameters in a

hierarchical model can cause big problems: ‘‘generally

speaking, if identifiability problems go undiagnosed,

inferences on these model parameters and possibly

others can be misleading.’’ For example, they warn that

Ver Hoef and Frost’s harbor seal data do not provide

enough information to estimate both within population

variation (as measured by the negative binomial

dispersion parameter) and among-population variation.

Unfortunately, it is hard to give a general prescription

for avoiding weakly unidentifiable parameters, except to

stress common sense again. If it is hard to imagine how

one could in principle distinguish between two sources

of variation—if different combinations of (say) between-

year variation and overdispersion would not lead to

markedly different patterns—then they may well be

unidentifiable.

3) Increase model complexity, and select models,

cautiously. Hierarchical models tempt users to include

lots of detail. Bayesian MCMC approaches will often

give an answer on problems where frequentist models

would crash, but the answers may be nonsensical

because of convergence problems. Starting chains in

different locations and examining graphical and quanti-

tative summaries of chain movement are supposed to

identify these problems, but experienced MCMC users

know that one can still be fooled by a sampler that gets

stuck for many steps. Bayesian statisticians tend to be

less concerned with simplifying models than frequent-

ists, in large part because they recognize that small

effects are never really zero, and perhaps because they

have traditionally worked in data-rich areas such as the

social sciences. However, they still recognize the

importance of parsimony: too-complex models will

predict future observations poorly, because the model

has been forced to fit the noise in the system as well as

the underlying regularities.

Model selection tools are similarly tempting.

Thoughtless, automated model selection rules always

lead to trouble: newer approaches such as AIC are a big

improvement over older stepwise techniques (Whitting-

ham et al. 2006), but are still subject to abuse (Guthery

et al. 2005). Bayesians have developed some tools, such

as DIC (Spiegelhalter et al. 2002) or posterior predictive

loss (Gelfand and Ghosh 1998), to estimate appropriate

model complexity, but because of problems with

convergence and identifiability, it is much better to use

self-discipline to limit model complexity to a level that

one can expect to fit from data. (I predict that ecologists

will rapidly begin to misuse deviance information

criterion (DIC), since it is computed automatically by

WinBUGS and is the easiest method of Bayesian

hierarchical model selection.)

4) Calibrate expectations. Without using model

selection tools, how can one know how much model

complexity one can expect to fit from a given data set?

Some classic rules of thumb, such as needing on the

order of 10 points per experimental unit (Gotelli and

Ellison 2004), or 10–20 points per estimated parameter

(Harrell 2001, Jackson 2003), or needing at least five to

six units to estimate variances, can be helpful. Reading

peer-reviewed papers, and paying careful attention to

the size of the data sets, their noisiness, and the

complexity of the models, is also useful. If all the

examples you can find that use the kinds of models you

want to fit are working with much larger data sets that

yours, watch out! McCarthy (2007) gives a variety of

examples using MCMC and Bayesian methods for

ecological inference on small, noisy data sets, but his

models are very simple and many of them use strongly

informative priors to fill in holes in the data.

5) Use simulated data to test models and develop

intuition. Perhaps the best way to determine whether a

model is appropriately complex for the data is to

simulate the model with known parameter values and

reasonable (even optimistic) sample sizes, and then to try

to fit the model to the simulated ‘‘data.’’ Simulated data

are a best-case scenario: we know in this case that the

data match the model exactly; we know whether our

fitted parameters are close to the true values; and we can

simulate arbitrarily large data sets, or arbitrarily many

smaller data sets, to assess the variability of the

estimates (or the statistical power for rejecting a null

hypothesis, or the validity of our confidence intervals)

and determine whether the model could work given a

large enough data set. Software bugs are common in

complex hierarchical models; data simulation is a way to

avoid the intense frustration caused by attempting to

debug a model and estimate parameters at the same

time. Simulating data is also a great way to gain general

understanding and intuition about a model. Simulating

data may seem daunting, but in fact a well-defined

statistical model is very easy to simulate. Defining the

FORUM590Ecological Applications

Vol. 19, No. 3

statistical model specifies how to simulate it. For

example, to simulate a single value from the Ver Hoef

and Frost seal model, one could run the BUGS model

above with specified parameter values, or in the R

language, one could use the commands

theta½t�,�rnormðn¼1;mean¼mu;sd¼sigmaÞY½i;t�,�rpoisðn¼1;expðtheta½t�ÞÞ:

ð3Þ

These commands are similar to the original statistical

description (Eq. 1) and the BUGS model (Eq. 2). (The

parameterization of the normal distribution has changed

yet again, with variability expressed in terms of the

standard deviation rather than the variance or the

inverse variance, and we use the R commands rnorm

and rpois to generate normal and Poisson random

deviates rather than dnorm and dpois as in BUGS.)

BENDING RULES

Expert statisticians often bend statistical rules. For

example, Cressie et al. spend a whole section describing

the need to think carefully about experimental design,

but they then admit that Ver Hoef and Frost’s seal study

was not randomized due to logistical constraints. How

does one decide that this nonrandom sampling design is

acceptable? (Since Ver Hoef and Frost’s goal is assess

trends only within this one metapopulation, they are at

least not making the mistake of extrapolating to a larger

population, but how do we know that this ‘‘convenience

sample’’ is not biased?) They say that one should try

fitting the model with a series of different prior

distributions to compare inferences: if the prior repre-

sents previous knowledge, how can one play with the

prior in this way? How did Ver Hoef and Frost decide

which factors to incorporate in their model and which

(such as spatial or temporal autocorrelation) to leave

out? Other such examples abound. For example, Hil-

born and Mangel (1997) use AIC to conclude that a

simple model provides the best fit to a set of fisheries

data, but they proceed with a more complex model

because they feel that the simple model underestimates

the variability in the system.

While researchers have developed non-Bayesian ap-

proaches to hierarchical models (de Valpine 2003,

Ionides et al. 2006, Lele et al. 2007), these tools are

typically either experimental or much harder to imple-

ment than Bayesian hierarchical models, and so the vast

majority of hierarchical models are Bayesian. Despite

Dennis’s (1996) claims, Bayesians are not as a whole

sloppy relativists; however, they do in general see the

world more in shades of gray than frequentists. Rather

than rejecting or failing to reject null hypotheses, they

compute posterior probabilities. Rather than use formal

tests to assess the validity of model assumptions and

goodness of fit, they look at graphical summaries

(Gelman and Hill [2006]; even frequentist statisticians

would prefer a good graphical summary to a thoughtless

test of a null hypothesis). Other, non-philosophical

reasons that Bayesians, and hierarchical modelers in

general, tend to be more flexible, are first that they have

often come from areas like environmental and social

sciences where huge (albeit noisy) data sets are available,

and hence they have had to worry a bit less about P

values (everything is significant with a large enough data

set) and strict model selection criteria; and second, they

have traditionally been more statistically sophisticated

than average users of statistics (and hence less likely to

cling to black-and-white rules).

My point here is not to say that any of these practices

is wrong, but rather to point out that despite many

ecologists’ desire for a completely objective way to make

inferences from data, the new world of hierarchical

models will require a lot of judgment calls. The

pendulum swings continually between researchers who

call for greater rigor (such as Hurlbert’s [1984] classic

condemnation of pseudoreplication) and those who

point out that undue rigor, or rigidity, runs the risk of

answering the wrong questions with high precision

(Oksanen 2001). Since hierarchical models are so

flexible, ecologists will have to make a priori decisions

about which parameters and processes to exclude from

their models. Giving up the illusion of the complete

objectivity of procedure-based statistics and dealing with

the complexities of modern statistics will be at least as

hard as learning the nuts and bolts of hierarchical

modeling.

ACKNOWLEDGMENTS

Nat Seavy and James Vonesh contributed useful commentsand ideas.

LITERATURE CITED

Berger, J. 2006. The case for objective Bayesian analysis.Bayesian Analysis 1:385–402.

Clark, J. S. 2007. Models for ecological data: an introduction.Princeton University Press, Princeton, New Jersey, USA.

Clark, J. S., A. E. Gelfand, and J. Samuel. 2006. Hierarchicalmodelling for the environmental sciences: statistical methods.Oxford University Press, Oxford, UK.

Cressie, N. A. C., C. A. Calder, J. S. Clark, J. M. Ver Hoef, andC. K. Wikle. 2009. Accounting for uncertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19:553–570.

Dennis, B. 1996. Discussion: Should ecologists becomeBayesians? Ecological Applications 226:1095–1103.

de Valpine, P. 2003. Better inferences from population-dynamics experiments using Monte Carlo state-space like-lihood methods. Ecology 84:3064–3077.

Edwards, D. 1996. Comment: The first data analysis should bejournalistic. Ecological Applications 6:1090–1094.

Gelfand, A. E., and S. K. Ghosh. 1998. Model choice: aposterior predictive loss approach. Biometrika 85:1–11.

Gelman, A., and J. Hill. 2006. Data analysis using regressionand multilevel/hierarchical models. Cambridge UniversityPress, Cambridge, UK.

Gotelli, N. J., and A. M. Ellison. 2004. A primer of ecologicalstatistics. Sinauer, Sunderland, Massachusetts, USA.

Guthery, F. S., L. A. Brennan, M. J. Peterson, and J. J. Lusk.2005. Invited paper: information theory in wildlife science:critique and viewpoint. Journal of Wildlife Management2369:457–465.

April 2009 591HIERARCHICAL MODELS IN ECOLOGY

Harrell, F. E., Jr. 2001. Regression modeling strategies.Springer, New York, New York, USA.

Hilborn, R., and M. Mangel. 1997. The ecological detective:

confronting models with data. Princeton University Press,Princeton, New Jersey, USA.

Hurlbert, S. 1984. Pseudoreplication and the design ofecological field experiments. Ecological Monographs 54:187–211.

Ionides, E. L., C. Breto, and A. A. King. 2006. Inference fornonlinear dynamical systems. Proceedings of the National

Academy of Sciences (USA) 103:18438–18443.

Jackson, D. L. 2003. Revisiting sample size and number ofparameter estimates: some support for the N : q hypothesis.

Structural Equation Modeling 10:128–141.

Kerman, J. 2007. Umacs: universal Markov chain sampler. Rpackage version 0.924. hhttp://cran.r-project.org/web/

packages/Umacs/iLele, S. R., B. Dennis, and F. Lutscher. 2007. Data cloning:

easy maximum likelihood estimation for complex ecologicalmodels using Bayesian Markov chain Monte Carlo methods.

Ecology Letters 10:551–563.

Martin, A. D., K. M. Quinn, and J. H. Park. 2008.MCMCpack: Markov chain Monte Carlo (MCMC) package.R package version 0.9-4. hhttp://mcmcpack.wustl.edui

McCarthy, M. 2007. Bayesian methods for ecology. CambridgeUniversity Press, Cambridge, UK.

Oksanen, L. 2001. Logic of experiments in ecology: ispseudoreplication a pseudoissue? Oikos 94:27–38.

Okuyama, T., and B. M. Bolker. 2005. Combining genetic andecological data to estimate sea turtle origins. EcologicalApplications 15:315–325.

Spiegelhalter, D. J., N. Best, B. P. Carlin, and A. Van derLinde. 2002. Bayesian measures of model complexity and fit.Journal of the Royal Statistical Society B 64:583–640.

Ver Hoef, J. M., and K. Frost. 2003. A Bayesian hierarchicalmodel for monitoring harbor seal changes in Prince WilliamSound, Alaska. Environmental and Ecological Statistics 10:201–209.

Whittingham, M. J., P. A. Stephens, R. B. Bradbury, and R. P.Freckleton. 2006. Why do we still use stepwise modelling inecology and behaviour? Journal of Animal Ecology 2675:1182–1189.

Woodworth, G. G. 2004. Biostatistics: a Bayesian introduction.Wiley, Hoboken, New Jersey, USA.

________________________________

Ecological Applications, 19(3), 2009, pp. 592–596� 2009 by the Ecological Society of America

Preaching to the unconverted

MARIA URIARTE1

AND CHARLES B. YACKULIC

Department of Ecology, Evolution and Environmental Biology, Columbia University, 10th Floor Schermerhorn Extension,1200 Amsterdam Avenue, New York, New York 10027 USA

Rapid advances in computing in the past 20 years

have lead to an explosion in the development and

adoption of new statistical modeling tools (Gelman and

Hill 2006, Clark 2007, Bolker 2008, Cressie et al. 2009).

These innovations have occurred in parallel with a

tremendous increase in the availability of ecological

data. The latter has been fueled both by new tools that

have facilitated data collection and management efforts

(e.g., remote sensing, database management software,

and so on) and by increased ease of data sharing

through computers and the World Wide Web. The

impending implementation of the National Ecological

Observatory Network (NEON) will further boost data

availability. These rapid advances in the ability of

ecologists to collect data have not been matched by

application of modern statistical tools. Given the critical

questions ecology is facing (e.g., climate change, species

extinctions, spread of invasives, irreversible losses of

ecosystem services) and the benefits that can be gained

from connecting existing data to models in a sophisti-

cated inferential framework (Clark et al. 2001, Pielke

and Connant 2003), it is important to understand why

this mismatch exists. Such an understanding would

point to the issues that must be addressed if ecologists

are to make useful inferences from these new data and

tools and contribute in substantial ways to management

and decision making.

Encouraging the adoption of modern statistical

methods such as hierarchical Bayesian (HB) models

requires that we consider three distinct questions: (1)

What are the benefits of using these methods relative to

existing, widely used approaches? (2) What are the

barriers to their adoption? (3) What approaches would

be most effective in promoting their use? The first

question has to do with motivation, that is, why does

one build a complex statistical model? Like Cressie et al.

(2009) we argue that while the goal of a model may be

estimation, prediction, forecasting, explanation, or

simplification, the purpose of modeling is the synthesis

of information. However, HB methods are not the only

tools available for synthesis (Hobbs and Hilborn 2006).

So the question needs to be refined to address the

specific benefits to be derived from HB models relative

Manuscript received 19 April 2008; revised 16 May 2008;accepted 16 June 2008. Corresponding Editor: N. T. Hobbs. Forreprints of this Forum, see footnote 1, p. 551.

1 E-mail: [email protected]

FORUM592Ecological Applications

Vol. 19, No. 3

to more traditional statistical approaches vis-a-vis

specific user goals. The second question deals primarily

with education, which we believe to be the main barrier

to the widespread adoption of these methods. Lastly,

answers to the third question build on the answers of

questions 1 and 2 to propose a series of actions that

would lead to a wider use of HB methods in ecology.

1. What are the benefits to be derived from HB models

relative to other statistical tools?.—Statistical modeling

in general and HB modeling in particular, are powerful

means to synthesize diverse sources of information.

With respect to other statistical means of synthesis,

hierarchical models have the advantage of allowing us to

coherently model processes at multiple levels. Consider,

for example, how we might answer the question of the

extent to which 10 species growth rates differ and

whether differences between tree species in growth rates

are correlated to some species trait, ST. One option

might be to first fit separate models using growth data

for each individual species together with important

covariates (e.g., individual level measurements), and

then use the results to fit a regression of the mean

growth of each species versus their mean ST values.

Another option might be to fit all the data at once and

include the ST repeated for all individuals in the plot.

Although each of these approaches might work ad-

equately, consider now that you have 100 species with

unequal sample sizes. With hierarchical models we could

include predictors at both the species and individual

levels and allow for partial pooling to improve

inferences on rarer species in a way that does not ignore

the initial uncertainty in the species growth estimates

when estimating the effect of ST across species.

Although the above statistical model could be fit using

non-Bayesian hierarchical models, HB becomes a

superior choice as we try to incorporate more of our

understanding of a process into a model. Returning to

the example above, consider the case in which there is

spatial autocorrelation between individuals sampled in

the same area and we realized that growth was measured

with error. Both are real concerns that we might

typically ignore or deal with in some ad hoc way;

however in a HB framework these sources of error could

easily be included an estimated as long as we had an

adequate data set.

In addition to their value for synthesis, and of far

more pragmatic significance, is the value of HB as a tool

for inference, particularly through the process of model

checking. The majority of ecologists seek to use data to

infer which processes are key in structuring populations,

communities and ecosystems. Inference is at the heart of

our discipline and therefore attaining the statistical

literacy necessary to use HB models can be extremely

rewarding, since such models allow us to incorporate the

complex variance structures inherent in most biological

systems. By working with simulated data derived from

HB models, rather than simple point estimates (with

associated confidence intervals), we can capture infer-

ential uncertainty and propagate it into predictions in a

straightforward manner (Gelman and Hill 2006). The

ability to not only make predictions from models but

also to quantify the uncertainty in our predictions, is

imperative for providing sound scientific advice for

management and policy decision makers.

For biologists interested solely in basic, rather than

applied questions, prediction serves as an important tool

for inference. If we approach a problem with an open

mind and the understanding that models are always an

approximation of reality, then comparison of the actual

data to replicate observations drawn from the posterior

predictive distribution (i.e., simulated observations

based on our model) becomes a learning exercise rather

than an effort to formalize what we already know or

believe. Although Cressie et al. (2009) argue that model

checking is necessary, but tedious, we see it not only as

the key to inference but also as one of the strongest

selling points of HB models. Prediction using traditional

statistical tools is limited, allowing only for a very

limited representation of the true complexity of eco-

logical data. In this context, model checking is an

opportunity to truly understand what the models are

saying, learn which parts of reality are not captured

adequately and suggest future steps. In particular, if

simulated data sets do not match the original data sets

adequately it leads directly to further model develop-

ment, reexamination of our interpretation of prior

studies, or alterations in experimental design for

collection of additional data (Fig. 1).

That model development typically follows model

checking illustrates that actual modeling of complex

data sets is typically an iterative process. Multiple

simpler models are fitted before attempts at the full

hierarchical model that we may have had in mind all

along or that may have evolved as we critically evaluate

the process we are trying to model and better understand

the data. Published studies typically emphasized final

models but understanding the iterative process of model

checking and model development is a key to demystify-

ing modeling to an audience of beginners, who are often

supplied only with unfamiliar technical descriptions of

models in the methods and little discussion of model fit

or misfit. Although standard statistical models caution

against extensive model checking because it can lead to

overfitting (data dredging), in a simulation framework

checks are means of understanding the limits of the

models’ applicability in realistic replications rather than

a reason for accepting or rejecting a model.

From a conceptual perspective, HB models offer a

consistent framework that allows the user to apply a

large, flexible number of models with complex variance

structures (e.g., repeated-measures models, time series

analysis, simultaneous consideration of observation and

process error, and so on). This is important not only

because we can tackle more complex problems but also

because it offers a way to educate students and

practitioners in a more self-consistent and coherent

April 2009 593HIERARCHICAL MODELS IN ECOLOGY

approach to statistical analysis that gets away from what

has termed a ‘‘field-key’’ approach to statistics, where

students collect statistical tools and techniques but fail

to see any connections among them at a deep level

(Clark 2005, Hobbs and Hilborn 2006).

2. What are the main barriers to the adoption of HB

methods?.—Using HB methods is not easy. There are

considerable conceptual and computational barriers to

overcome. Conceptually, students must move from a

descriptive, test-based statistical framework to an

inferential, estimation-based, complex one. Learning to

use HB methods requires a larger initial investment to

gain a holistic understanding of statistical inference, as

opposed to the short term solutions of finding a test for

the question at hand or restricting oneself to questions

that can be answered with the tests we already know.

HB methods may not always be the best tool for

answering a particular question, and often simpler

methods may be adequate. Nonetheless, learning HB

opens up the possibility of addressing a range of

previously intractable questions that more accurately

encompass the complexity of biological systems. As

ecologists accrue larger datasets, much of it based on

remotely gathered data with multiple sources of error,

the potential benefits of adopting more complex models

increase, moving the curve in Fig. 1 further to the right.

Often, students will not see the need for the initial

investment in learning HB methods because they lack an

understanding of its potential benefits and they perceive

modeling as a skill rather than as a tool that anyone can

pick up. Thus, ignorance begets indifference, or worst,

fear. If they see a paper in say, Ecological Applications

exhorting the benefits of HB models, they are likely to

turn the page and dismiss it as just another modeling

paper or simply over their heads. Even if a student or

practitioner is interested, the barriers may seem insur-

mountable without at least some knowledge of either

programming or advanced statistical methods.

Indeed, such knowledge is a prerequisite for learning

HB methods. One way to acquire these skills is through

formal graduate-level courses. Curricula that connect

models and quantitative thinking to important questions

in ecology have proven to both ignite students’ interest

in modeling and to convey the relevance and usefulness

FIG. 1. The model fitting process often consists of fitting progressively more complex models (e.g., A, then B) and/or trying andfailing at fitting more complex models (e.g., G, then F) and working backward until one finds a more simplified model that can befit with the data available (e.g., E). The exact location of the cutoff between E and F will depend on the nature of the data at hand.Knowing which part of reality to allow back into the model by relaxing assumptions or partitioning uncertainty is dependent bothon an understanding of the ecological question, the data, and one’s statistical literacy. As model complexity increases, one can moreclosely approximate reality, include more substantial outside (prior) knowledge/intuition, and gain more confidence in the modeloutput and associated uncertainty; however, added complexity only helps if ecological understanding is properly translated into themodel structure (the daggers in the figure indicate this caveat).

In the example, Cressie et al. (2009) address a number of models of differing complexity. Model A might correspond to a simplelinear regression of numbers vs. time. Such unrealistically simplified models could potentially lead to estimates with tight confidenceintervals and low P values, but unreliable inference. Model B might correspond to a generalized linear model with Poissondistributed errors, whereas model C might correspond to the simplest model considered by Cressie et al. (2009), a generalized linearmixed model with multiple explanatory variables. In this context, model C has the advantage that it is no longer assumed that allsites are the same, something we know is false. Partitioning uncertainty into all its potential components and adding site-specificparameters may lead to a model F that cannot be fit with the data at hand, while adding in the assumption that site parameters arerelated and come from a distribution may result in model D.

FORUM594Ecological Applications

Vol. 19, No. 3

of models. However, many ecology and evolution

programs still rely on statistics departments to train

their students and only a few offer advanced statistical

methods courses within their department. Farming out

ecology students to be trained in statistics departments is

far from ideal because courses are likely to be developed

to address the needs of statistics students rather than

those of other areas of science. If students fail to see the

relevance of the methods to their own discipline,

motivation will decline.

Although there are a number of open, short-term

courses available (Duke University summer course in

ecological forecasting, one-day workshop in Bayesian

methods at the Ecological Society of America annual

meetings), these offerings are limited, unpredictable, and

costly. Moreover, short courses education in HB models

often requires not only some programming skills but a

complete conceptual overhaul of the students’ existing

conceptual statistical framework. Although short

courses may offer an entry into HB models, the tension

between a focus on tools (e.g., WinBUGS) and concepts

is very real when time is limited.

A second way to learn HB is through self-teaching.

Although a few books have appeared in the last couple

of years that make self-teaching possible (Woodworth

2004, Gelman and Hill 2006, Clark 2007, McCarthy

2007), it is still a daunting task to learn these methods on

your own without a support network. Fortunately,

cyber-communities are becoming an increasingly im-

portant form of scientific exchange, and considerable

progress can be made in this way (e.g., contributions and

exchanges around the development and use of R and

WinBUGS Statistical freeware). Auto-didactic ap-

proaches are powerful means to learn but they often

leave behind major lacunas in knowledge because they

tend to focus on the mechanics of carrying out complex

statistical analysis with little attention paid to the

foundations of underlying statistical inference (e.g.,

Taper and Lele 2004).

Another barrier to the adoption of HB is the lack of

consensus in many of the details of implementation. HB

methods are relatively new and there seems to be a lack

of consensus on what are the best non-informative

priors to use, how best to assess convergence, the utility

of deviance information criterion (DIC), etc. This is a

major barrier to biologists who are trying to learn and

implement these methods, since there is often no clear

path to follow. Many ecologists will choose those

methods that they know well despite their shortcomings.

3. What approaches would be most effective in

promoting their use?.—Although there are a small

number of graduate ecology programs that train some

students in modern statistics including HB methods, an

impediment to widespread training teaching in these

methods is the availability of ecology faculty with

statistical background sufficient to offer such courses.

Faculty members need efficient and relevant ways to get

training in modern statistical modeling and the neces-

sary tools and materials to teach them effectively. Given

the time demands placed on faculty, self-education

approaches may be unrealistic. One potential solution

is to offer one-semester sabbatical leaves that interested

faculty could use to attend existing courses at univer-

sities where such courses are offered. Better yet, these

leaves could be structured around the development of

intensive courses that brought together a small group of

expert teachers and student-faculty. The advantages of

this approach are threefold. First, the burden of

teaching could be shared among a small group of

experts. Second, the students would be exposed to a

variety of viewpoints. Third, participating faculty would

gain not only technical and conceptual skills but also a

support network to carry the newly acquired skills back

to their home institutions. This approach would

probably work best with faculty who are already

teaching statistics or use modeling in their research, or

with postdoctoral researchers who have the time and

motivation to learn and use the methods. Costs could be

shared by interested institutions and funding agencies.

In institutions that lack faculty trained in modern

statistical methods, ecology departments could also

work with faculty in the statistics department to develop

advanced courses or at least to discuss statistical issues

and problems as they arise. The advantage of this

approach is that students would be exposed to both the

rigor of statistics and disciplinary applications. The

shortcomings are generating faculty interest and the

considerable investments required to develop a course

with multiple instructors from different disciplines and

departments. Existing or new collaborations between

statisticians and ecologists within the same institution

could be leveraged to this end. Educators in ecology and

other fields that depend on statistics departments for

introductory courses could also initiate a dialog with

their statistical colleagues about restructuring ‘‘service’’

courses to cover basic concepts that are at the heart of

modern statistical methods such as distributional theory.

University administrators could also be approached to

offer faculty incentives for the development of courses

that cross disciplinary boundaries.

Students can also act as catalysts in the adoption of

modern statistics in ecology. Although we caution

against the perils of unabashedly using graduate

students as a means to improve existing programs, both

students and faculty have much to gain from judicious

small-scale efforts. Graduate students are often looking

for teaching opportunities that give them some experi-

ence in curriculum development and allow flexible

didactic approaches. At the same time, faculty members

are also searching for means to increase students’

engagement. One way to address the goals of these

two groups is to allow graduate students to structure

parts of existing course or to have them offer workshops

that provide an introduction to the tools and techniques

that will facilitate self-teaching for other students. For

instance, a short course in R, or structuring labs in

April 2009 595HIERARCHICAL MODELS IN ECOLOGY

existing statistics courses in R software rather than acommercial package, will provide students with some

familiarity with programming and open the door to alarge cyber community with which they can engage. Thisapproach would require some thought on the part of the

faculty and students but could potentially be verypowerful because students readily accept new knowledgeand methods from their peers.

Cyber courses can also be a means to bring statisticalliteracy to ecologists. This approach could make anumber of existing graduate courses in modern statis-

tical methods accessible to ecologists. These coursesinclude among others Ecological Models and Data atthe University of Florida, Ecological Theory and Dataat Duke University, Modeling for Conservation of

Populations at the University of Washington, andSystems Ecology at Colorado State University. Withrelatively minor investments, these courses could be

broadcasted to other graduate program in the UnitedStates and abroad. Although interactions betweenstudents and teaching faculty would be limited, this

approach would provide students with a foundation topursue further education in statistics either through self-teaching or collaboration with statistics faculty at their

home institutions. To encourage use and discussion, thecyber courses could be structured around student–faculty groups at the receiving institutions.Ultimately the widespread adoption of modern

statistical methods will require a mix of approaches.What makes sense for individual institutions will dependon the availability of faculty and on motivations to

develop offerings in this area. Funding agencies can help

by providing incentives to institutions and individual

faculty. To the degree that faculty and students are

interested and willing, statistical literacy can be devel-

oped.

ACKNOWLEDGMENTS

We thank Liza Comita for constructive comments.

LITERATURE CITED

Bolker, B. M. 2008. Ecological models and data in R. PrincetonUniversity Press, Princeton, New Jersey, USA.

Clark, J. S. 2005. Why environmental scientists are becomingBayesians. Ecology Letters 8:2–14.

Clark, J. S. 2007. Models for ecological data: an introduction.Princeton University Press, Princeton, New Jersey, USA.

Clark, J. S., et al. 2001. Ecological forecasts: an emergingimperative. Science 293:657–660.

Cressie, N., C. A. Calder, J. S. Clark, J. M. Ver Hoef, and C. K.Wikle. 2009. Accounting for undertainty in ecologicalanalysis: the strengths and limitations of hierarchicalstatistical analysis. Ecological Applications 19:553–570.

Gelman, A., and J. Hill. 2006. Data analysis using regressionand multilevel/hierarchical models. Cambridge UniversityPress, New York, New York, USA.

Hobbs, N. T., and R. Hilborn. 2006. Alternatives to statisticalhypothesis testing in ecology: a guide to self teaching.Ecology 16:5–19.

McCarthy, M. A. 2007. Bayesian methods for ecology. Cam-bridge University Press, Cambridge, UK.

Pielke, R. A., and R. T. Connant. 2003. Best practices inpredictions for decision-making: lessons from the atmos-pheric and earth sciences. Ecology 84:1351–1358.

Taper, M. L., and S. R. Lele. 2004. The nature of scientificevidence: statistical, philosophical, and empirical consider-ations. University of Chicago Press, Chicago, Illinois, USA.

Woodworth, G. G. 2004. Biostatistics: a Bayesian introduction.Wiley, Hoboken, New Jersey, USA.

FORUM596Ecological Applications

Vol. 19, No. 3